<<

LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Lev Vasilievich Shubnikov „1901–1937…: On the centennial of his birth ͓DOI: 10.1063/1.1414585͔

A number of brilliant studies were done at this cryogen- ics laboratory, among them: — direct experimental proof ͑and independent of the Meissner and Ochsenfeld results͒ of the ideal diamagnetism of pure superconductors ͑Yu. N. Rjabinin and L. W. Schub- nikow, Nature 134, 286 ͑1934͒͒; — observation of an antiferromagnetic phase transition ͑jump of specific heat͒ in layered transition-metal chlorides ͑O. N. Trapeznikowa and L. W. Schubnikow, Nature 134, 378 ͑1934͒͒; — studies of the phase diagrams and viscosity of liquid mixtures of nitrogen, oxygen, carbon monoxide, methane, argon, and ethylene ͑jointly with O. N. Trapeznikova and N. S. Rudenko͒; — the experimental discovery of type-II superconduct- ors ͑Yu. N. Rjabinin and L. W. Schubnikow, Nature 135, 581 ͑1935͒; L. W. Schubnikow, W. I. Chotkewitsch, G. D.

September 29, 2001 marks the 100th anniversary of the birth of Lev Vasilievich Shubnikov, one of the most out- standing experimental physicists of the twentieth century. Shubnikov was the founder and director of the first cryogen- ics laboratory in the USSR, and his pioneering work laid the foundation for many extremely important fields in modern condensed-matter physics. In terms of the quantity and the level of the results obtained in various fields of physics, he can be placed in the same rank of such giants of experimen- tal physics as Faraday, Kelvin, and Kamerlingh Onnes. Shubnikov’s scientific career is linked to three cities: Petrograd ͑1923–1926͒ — the creation of a new method of growing single crys- tals ͑the Obreimov–Shubnikov method, Z. Phys. 25,31 ͑1924͒͒; — the creation of an optical method of studying plastic deformations in crystals ͑I. W. Obreimow and L. W. Schub- nikow, Z. Phys. 41, 907 ͑1927͒͒, Leiden ͑1926–1930͒ — discovery of magnetoresistance oscillations at low temperatures ͑the Shubnikov–de Haas effect, Leiden Com- mun. 207 a–d; 210 a,b; Nature 126, 500 ͑1930͒͒, and Kharkov ͑1930–1937͒. Here, at the Ukrainian Physicotechnical Institute, Shub- nikov organized the first cryogenics laboratory in the Soviet ͑ ͒ Union and the fourth in the world , which he directed until FIG. 1. Cover of the Leiden laboratory journal containing the first publica- 1937. tion of the Shubnikov–de Haas effect.

1063-777X/2001/27(9–10)/2/$20.00689 © 2001 American Institute of Physics 690 Low Temp. Phys. 27 (99–10), September–October 2001

— experimental observation of an intermediate state of superconductors in magnetic field ͑L. W. Schubnikow and I. Nakhutin, Nature 139, 589 ͑1937͒͒; — identification of the magnetic character of a phase transition ͑the kink in the temperature dependence of the magnetic susceptibility in transition-metal chlorides͒͑L. W. Schubnikow and S. S. Schalyt, Phys. Z. Sowjetunion 11, 566 ͑1937͒͒; — a study of the thermal conductivity of solid helium ͑jointly with K. A. Kikoin ͑1936–1937͒͒; — a study of a phase transition in solid methane under pressure ͑jointly with O. N. Trapeznikov and G. A. Milyutin, FIG. 2. Handwritten notes of L. V. Shubnikov in his laboratory journal on Nature 144, 632 ͑1939͒͒. ͑ obtaining the first batch of liquid hydrogen from the archive of B. G. In 1937 L. V. Shubnikov became a victim of Stalinist Lazarev͒. 12/XI 31. First test of large liquid hydrogen machine . . . terror. 14/XI 31. Obtained hydrogen in the amount of ca. 3 liters . . . In terms of the number of brilliant scientific results and extremely fruitful new directions of research begun and fos- tered in its six years of existence, Shubnikov’s laboratory is apparently without equal in the world. The creative explosion that marked the period 1934– 1937 attested to Shubnikov’s creative energy and scientific potential of the highest sort, which had only begun to be Schepelew, and Yu. N. Rjabinin, Phys. Z. Sowjetunion 10, realized in the lifetime of this outstanding physicist of the ͑ ͒͒ 165 1936 ; twentieth century. — measurement of the magnetic moment of the proton Yu. A. Fre man ͑B. G. Lasarew and L. W. Schubnikow, Phys. Z. Sowjetunion  10,117͑1936͒; ibid. 11, 445 ͑1937͒͒; — optical studies of liquid helium II ͑A. K. Kikoin and This anniversary issue of the journal Low Temperature L. W. Schubnikow, Nature 138, 641 ͑1936͒͒; Physics is devoted to the memory of Lev Vasilievich Shub- — a study of neutron absorption at low temperatures in nikov. It includes articles on the physics of quantum mag- hydrogen, boron, silver, and cadmium ͑jointly with V. Fo- netic oscillation phenomena, superconductivity, magnetism, min, F. Houtermans, I. V. Kurchatov, A. I. LeŽpunskiŽ,L.B. and cryocrystals — fields of study whose creation owes Rusinov, and G. Ya. Shchepkin, Nature 138, 326 ͑1936͒; much to the contributions made by Shubnikov. The Editorial ibid. 138, 505 ͑1936͒͒; Board of this journal extends its profound gratitude to all of — a study of the destruction of superconductivity by the authors who contributed to this issue. electric current and magnetic field ͑L. W. Schubnikow and Editorial Board N. E. Alexeyevskii, Nature 138, 545 ͑1936͒; ibid. 138, 804 ͑1937͒͒; Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

The Shubnikov–de Haas effect and high pressure E. S. Itskevich*

L. F. Vereshchagin Institute of High Pressure Physics, Russian Academy of Sciences, 142190 Troitsk, Moscow District, Russia ͑Submitted March 14, 2001͒ Fiz. Nizk. Temp. 27, 941–944 ͑September–October 2001͒ ͓DOI: 10.1063/1.1401175͔

Dedicated to the memory of L. V. Shubnikov and O. N. Trapeznikova

Lev Vasilievich Shubnikov, who was born 100 years ago noted that the theory of quantum oscillation effects was cre- at the dawn of the twentieth century ͑September 29, 1901͒, ated in the 1960s by I. M. Lifshits and A. M. Kosevich. would play a leading role in the development of quantum My acquaintance with the life and undertakings of L. V. physics of the solid state. A century has also passed since the Shubnikov grew out of a rather close acquaintance with O. birth of his wife and co-worker, Olga Nikolaevna Trapezni- N. Trapeznikova. It began in the 1960s during joint research kova, who was an active participant in much of Shubnikov’s on the specific heat of highly anisotropic solids and contin- work. The field of solid-state physics, which arose at the end ued for many years. Shubnikov and Trapeznikova’s son, of the first quarter century, underlies all of the achievements Misha Shubnikov, who is also a physicist, worked for many of electronics, the principal technological base of the infor- years in Shalyt’s laboratory at the Leningrad Physicotechni- mation revolution of the twentieth century. cal Institute and did research on the properties of semicon- Quantum oscillations of the electrical resistance in metal ductors under hydrostatic pressure, continuing the research single crystals in a magnetic field — the so-called on semiconductors under pressure which was begun there in Shubnikov–de Haas ͑SdH͒ effect — were discovered by 1965. The first paper was on a study of the influence of Shubnikov jointly with Prof. de Haas, his scientific supervi- pressure on the magneto-phonon oscillations in n-InSb ͑S. S. sor at Leiden, Holland ͑1930͒. Perfect single crystals of bis- Shalyt, E. S. Itskevich, and co-workers͒. muth, which made it possible to observe the SdH effect, were My acquaintance with Olga Trapeznikova was a multi- grown by Shubnikov as the first task assigned by de Haas to faceted one. Her long life ͑she died in 1997͒ was filled with the young scientist who had come to him for practical train- tragic circumstances. After the arrest and execution of her ing. Later these same crystals were used by de Haas and van husband Shubnikov in 1937 and her trials during the Second Alphen in their discovery of analogous oscillations of the World War, Trapeznikova found the strength to continue her magnetic susceptibility ͑the de Haas–van Alphen ͑dHvA͒ ef- scientific work at St. Petersburg University and to raise her fect͒. Both types of oscillations can be used to obtain experi- son. She was the perfect example of the St. Petersburg intel- mental information about the energy spectrum of the charge ligentsia, incorporating all the intellectual values of the early carriers in metals and semiconductors. twentieth century. Not only was she widely educated, she Before turning to the topic stated in the title of this ar- had a love of people and displayed a sympathy that was rare ticle, I would like to give some information about Shubnikov in those times. himself. He founded the first cryogenics laboratory in the It was also my good fortune to be associated ͑including Soviet Union, at the Kharkov Physicotechnical Institute in scientific collaboration͒ with other students of Shubnikov’s 1931. He did his practical training in Leiden ͑1926–1930͒ in who would gain lasting recognition in science: N. E. Alek- the laboratory of Kamerlingh Onnes, who, as you know, was seevskiŽ, B. G. Lazarev, and Yu. N. Ryabinin, in addition to the first to obtain liquid helium. Shubnikov mastered the art Trapeznikova and Shalyt. These were very productive and of experimentation with liquid helium and therefore was able pleasant associations. The most important contact was with in a comparatively short time to set up facilities at Kharkov the founder and director of the Institute of High Pressure for obtaining ‘‘all’’ the liquid gases — nitrogen, hydrogen, Physics of the Academy of Sciences, Leonid Fedorovich and helium, making it possible to do research on a broad Vereshchagin, whose scientific career in physics also started spectrum of low-temperature problems. In particular, signifi- at Shubnikov’s laboratory ͑1934–1939͒ with research on the cant work was done on superconductivity. Shubnikov’s ex- magnetic susceptibility of alloys. Although Vereshchagin’s periments gave food for thought to some of the greatest theo- further activities were devoted to high-pressure research, he rists of our country. His work with Trapeznikova on remembered his younger years in low-temperature studies. In measurement of the low-temperature specific heat of 1958, when forming the Institute of High Pressure Physics transition-metal chlorides and with S. S. Shalyt on measure- ͑IHPP͒ with the main task of mastering the synthesis of dia- ment of their magnetic susceptibility led to the discovery of monds at high pressure, he invited me, as a specialist in transitions to an antiferromagnetic state, the theory of which low-temperature physics and a representative of the Institute was created by L. D. Landau. Skipping ahead, it may be of Physical Problems, to effect another ‘‘synthesis’’ — to

1063-777X/2001/27(9–10)/4/$20.00691 © 2001 American Institute of Physics 692 Low Temp. Phys. 27 (99–10), September–October 2001 E. S. Itskevich combine high pressures and low temperatures in the study of the fundamental properties of solids. The field begun by Shubnikov was branching and developing.

THE SdH AND dHvA EFFECTS We shall not distinguish between these two quantum ef- fects, since in practice they give the same information about the object of study in different kinds of measurements. They arise as a result of the quantization of the energy levels of the conduction electrons in a magnetic field ͑Landau quantiza- tion͒ and are inherent to metals and semiconductors. These effects were discovered thanks to the availability of helium temperatures and the creation of methods of obtaining per- fect single crystals of metals and semiconductors. Both effects involve oscillations of certain properties in a magnetic field at low temperatures, and the possibility of observing them is contingent upon the quality of the samples, low temperatures, and a high external magnetic field. The oscillation period in the inverse magnetic field is related to the area of the extremal ͑minimal or maximal͒ section of the Fermi surface by a plane perpendicular to the field direction, ⌬(1/H)ϭeh/cS, where S is the area of the extremal cross section of the Fermi surface, and e, c, and h are well-known global constants. By varying the field direc- FIG. 1. Formation of a new cavity ͑a͒, the breakoff of the neck ͑b͒, and the tion, one can find the shape of the Fermi surface. By mea- transition of a system of corrugated tubes to ellipsoids ͑c͒. suring the temperature dependence of the amplitude of the oscillations one can determine the carrier effective mass cor- responding to the given cross section of the Fermi surface, and the field dependence of this same amplitude can be used HIGH-PRESSURE CHAMBERS to assess the scattering of carriers on impurities, i.e., the Low-temperature hydrostatic pressure chambers suitable quality of a crystalline sample. The SdH and dHvA effects for oscillation experiments were constructed in 1962–1963 have proved to be the most convenient and reliable methods ͑E. S. Itskevich͒. These were piston-and-cylinder systems in of experimental determination of the Fermi surface. the which a pressure medium ͑usually a mixture of organic liquids͒ was used in the working volume to transmit the pres- ELECTRONIC–TOPOLOGICAL TRANSITIONS UNDER sure to the crystalline sample. The piston generating the pres- PRESSURE sure could be fixed at different degrees of compression of the The creation of a high-pressure chamber permitting in- medium. This technique of creating hydrostatic pressure in a vestigation of the Fermi surface at high hydrostatic pressures volume with practically no loss of pressure is fundamentally by means of the oscillation effects was inspired by a paper different from the ‘‘clamped cell’’ technique proposed by by I. M. Lifshits ͑1960͒ in which he predicted qualitative Chester and Johnson in 1953, in which volumeless anvils are changes in the Fermi surface of metals ͑a phase transition of fixed in position and hydrostatic conditions are not realized. 1 order 2 2). Lifshits’s idea can be easily understood. The ͑ ͒ Fermi surfaces of metals in momentum wave-vector space ELECTRONIC–TOPOLOGICAL TRANSITIONS IN CADMIUM can be represented in the most general form as corrugated tubes ͑straight or curved; see Fig. 1͒. In the first case the Electronic–topological ͑ET͒ transitions are phase transi- 1 extremal cross section of the tube can coincide with the tions of order 2 2) under pressure. The destruction of the boundaries ͑faces͒ of the Brillouin zone — the elementary connectedness of the corrugated tubes of the Fermi surface Wigner–Seitz cell in momentum space in an extended zone of metals at an ET transition can be observed not only by scheme. In the second case the tube can have the form of a means of oscillation effects but also from the changes of the corrugated ring or system of rings placed inside the cell or on trajectories of the current carriers in a magnetic field, i.e., its edges ͑Fig. 1͒. One feels an irresistable urge to compress from the angle dependence of the magnetoresistance in a the extremal cross section of the Fermi surface ͑the neck͒ high static magnetic field. until it vanishes and thus to obtain separated closed parts of Single-crystal cadmium samples have been investigated the Fermi surface. This can naturally be done by applying by both methods, and complete agreement was found for the pressure. In principle the opposite process — the joining of characteristics of the ET transitions observed ͑S. L. Bud’ko, ͒ disconnected parts of the Fermi surface into single ‘‘tubes’’ A. N. VoronovskiŽ, A. G. Gapotchenko, and E. S. Itskevich . by applying pressure — is also possible, usually in the hole Figure 2 shows the dHvA oscillations in cadmium, and Fig. 3 parts of the Fermi surface. shows the ET transitions that occur. Low Temp. Phys. 27 (99–10), September–October 2001 E. S. Itskevich 693

Fisher, 1967͒ in semiconductors, jointly with the Leningrad Physicotechnical Institute ͑LPTI͒ and the Physics Institute of the Academy of Sciences ͑FIAN͒ in tellurium, and jointly

with LPTI in n- and p-Bi2Te3. Successful searches for ET transitions by oscillation methods have also been carried out in other laboratories: at the Department of Low Temperature physics at Moscow State University ͑N. B. Brandt and co- workers͒ and at the Donetsk Physicotechnical Institute ͑A. A. Galkin and co-workers͒. Also noteworthy is the beautiful work of N. Ya. Minina ͑Moscow State University͒, who has discovered a whole series of ET transitions in Bi and BiSb alloys using methods she developed based on the uniaxial deformation of single crystals. Thus a new field of solid-state physics was created: the investigation of qualitative changes ͑transitions͒ in the elec- tronic spectra of metals and semiconductors. In addition to ET transitions, oscillation methods under pressure have been used to study changes of the Fermi sur- face in magnetically ordered metals — Fe, Ni, Co, and Cr ͑IHPP and Institute of General Physics of the Academy of Sciences, 1977–1979͒. Altogether, our group has published more than 30 papers on the study of oscillation effects under pressure. FIG. 2. De Haas–van Alphen oscillations in cadmium, due to new parts of 1 Since many laboratories in different countries have been the Fermi surface arising at phase transitions of order 2 2; a — frequency of doing research on the various changes of the Fermi surface ʈ ¯ ʈ oscillations, H ͓1120͔; b — frequency of oscillations, H ͓0001͔; before ͑ ͒ transition ͑1͒, after transition ͑2͒. of metals not only ET transitions by means of oscillation methods, the accumulated experimental data is sufficient to permit assessing the applicability of different models under conditions of hydrostatic compression. It has turned out that These results, together with analogous results of B. G. pseudopotential theory works rather well. Lazarev and co-workers, the experiments of N. B. Brandt Recently the study of electronic spectra under pressure and Ya. G. Ponomarev, and the theory of I. M. Lifshits, were by means of the SdH oscillations has been extended to two- registered as a Discovery ͑Certificate No. 238͒. All of this dimensional systems. For example, it is possible to measure ͑ demonstrates the direct through the participating scientists the magnetoresistance and Hall emf in fields up to 7 T at T ͒ and their followers and conceptual links to Shubnikov as the ϭ4.2 K under pressures up to 2.5 GPa for systems with founder of a new field of solid-state physics. The GaSb/InAs/GaSb quantum wells of various widths and with Shubnikov–de Haas effect is the principal tool used in these different types of interfaces. The pressure dependence of the studies. Using the SdH oscillations, our group has found ET tran- electron and hole concentrations has been determined from sitions in bismuth and Bi–Sb alloys ͑E. S. Itskevich, L. M. an analysis of the observed SdH oscillations and the depen- dence of the magnetoresistance tensor components in the framework of a model with two types of carriers. It has been shown that a transition from the semimetallic to the semicon- ductor type of conduction, accompanied by the onset of a region of negative magnetoresistance, occurs under pressure; this region persists up to pressures corresponding to a tran- sition to the insulating state. The existence of a region of negative magnetoresistance can be explained in the frame- work of the concept of weak localization in a system of two-dimensional electrons. Quantitative information on the concentration and mobility of the carriers in the system has been obtained, and also on the relationship between the elas- tic and inelastic relaxation times for quasi-two-dimensional FIG. 3. Schematic illustration of the Fermi surface of cadmium. Open hole electrons and the pressure dependence of these times ͑E. M. surface — the ‘‘monster’’ ͑second Brillouin zone͒. There are discontinuities Dizhur et al., IHPP, 1998͒. in the basal plane, and open directions do not form along the ͓112¯ 0͔ and ͓101¯ 0͔ axes. At the K points there are no ‘‘needles.’’ The dashed lines show In closing, it can be said that L. V. Shubnikov created a transitions in cadmium: necks (␦) and ‘‘needles’’ (␧) arise. research tool that can be used to study the behavior of the 694 Low Temp. Phys. 27 (99–10), September–October 2001 E. S. Itskevich conduction electrons in metals and semiconductors under lossal magnetoresistance manganates; these have come to be pressure. called ‘‘quantum phase transitions.’’ The further development of solid-state physics has led to the discovery of transitions from some states to other states, E-mail: [email protected] with practically no change in crystal structure, in new ob- jects, in particular high-Tc superconducting cuprates and co- Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Orientational effect in the magnetoresistance of organic conductors V. G. Peschanskiı˘*

B. Verkin Institute for Low Temperature Physics and Engineering, National Academy of Sciences of Ukraine, pr. Lenina 47, 61103 Kharkov, Ukraine; V. N. Karazin Kharkov National University, pl. Svobody 4, 61077 Kharkov, Ukraine R. Atalla

Physics Department of Bir-Zeit University, P.O. Box 28, Bir-Zeit, West Bank, Israel ͑Submitted April 6, 2001͒ Fiz. Nizk. Temp. 27, 945–951 ͑September–October 2001͒ The asymptotic behavior of the magnetoresistance of layered conductors of organic origin is discussed for different orientations of a high magnetic field H relative to the layers. It is shown that when current flows along the normal n to the layers, the amplitude of the Shubnikov–de Haas quantum oscillations and the smoothly varying part of the magnetoresistance increase sharply at certain angles ␪ between the vectors H and n. An experimental study of the orientational effect in a high magnetic field yields detailed information about the electron energy spectrum of quasi-two-dimensional conductors. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1401176͔

The oscillatory dependence of the resistance of rather racene halides, which have a layered structure with a pro- perfect samples of bismuth as a function of the inverse mag- nounced anisotropy of the electrical conduction: the conduc- netic field, discovered by Shubnikov and de Haas in Leiden tivity along the layers is significantly higher than that in the in 1930,1,2 permitted the development of a reliable spectro- direction transverse to the layers. Consequently, the organic scopic method3,4 that is still being used successfully to re- conductors of this family have a metallic type of conductiv- construct the Fermi surface, a fundamental characteristic of ity with a high resistance to current flowing along the normal the electron energy spectrum of degenerate conductors, from to the layers, apparently because of the sharp anisotropy of p onץ/(␧(pץexperimental data. the velocities of the conduction electrons vϭ ␧ ϭ␧ The Shubnikov–de Haas ͑SdH͒ effect and the oscillatory the Fermi surface (p) F , i.e., their energy dependence of the magnetization as a function of the inverse ϱ 5 anp magnetic field, discovered the same year, were long consid- ␧͑ ͒ϭ ␧ ͑ ͒ ͭ z ϩ␣ ͑ ͒ͮ ͑ ͒ p ͚ n px, py cos n px, py , 1 ered to be anomalous behavior of bismuth, which also has nϭ0 h other unusual properties. However, these oscillations at low ␣ ͑ ͒ϭϪ␣ ͑Ϫ Ϫ ͒ temperatures were later observed in practically all metals, n px ,py n px , py and Landau showed that these effects are due to the presence is weakly dependent on the momentum projection p ϭp"n of features of the density of states of the conduction electrons z 6 on the normal n to the layers, so that the maximum values of in a quantizing magnetic field. Quantum oscillation effects ␧ у the functions n(px , py) with n 1 on the Fermi surface, have turned out to be inherent to all degenerate conductors ϭ ␧ ϭ␩ ␧ An max n(px , py) n F , are much less than the Fermi en- and are observed when the distance between the quantized ergy ␧ . ⍀ F levels of the electron energy h exceeds the temperature- The Fermi surface of quasi-two-dimensional conductors related smearing of the Fermi distribution function of the is open, with a slight corrugation along the pz axis; the cor- charge carriers but is considerably less than the Fermi rugated plane may be rolled up into a cylinder whose base is energy. positioned in the unit cell of momentum space in such a way In 1988 SdH oscillations of the magnetoresistance of the that the Fermi surface of the layered conductors is a system ͑ ͒ ͑ ͒ organic conductors BEDT–TTF 2IBr2 and BEDT–TTF 2I3 of slightly corrugated cylinders or a system of planes slightly were observed at sufficiently low temperatures in a magnetic corrugated along the pz axis. Small closed cavities associated field H of several tens of tesla, when the gyration frequency with anomalously small groups charge carriers may also ⍀ of the electrons is significantly higher than their collision exist. frequency 1/␶.7–9 In the case of a discrete or discrete–continuous spectrum Even more unexpected was the observation of periodi- of electron energies in a magnetic field the features of their cally repeating narrow peaks in the dependence of the resis- density of states repeat periodically with variation of 1/H, tivity ␳ on the angle ␪ between the direction of the field and this is the cause of the quantum oscillation effects. The vector of the high magnetic field and the normal n to the presence of quantum magnetoresistance oscillations of large layers during passage of current transverse to the layers.7,8 amplitude in a magnetic field Hϭ(0,H sin ␪, H cos ␪)inthe By now the SdH quantum oscillation effect has been family of tetrathiafulvalene salts indicates the existence of a observed in many tetrathiafulvalene salts and tetraselenotet- rather large number of charge carriers at the Fermi energy

1063-777X/2001/27(9–10)/5/$20.00695 © 2001 American Institute of Physics 696 Low Temp. Phys. 27 (99–10), September–October 2001 V. G. Peschansky and R. Atalla

ץ which execute finite motion in a plane orthogonal to the eHh cos ␪ 1/2 ␧ϭ ϭ " Pˆ ϭiͩ ͪ , magnetic field, i.e., their orbits const, pH p H/H x c ␧␰ ϭconst are closed, and at least one sheet of the Fermi sur- face is a slightly corrugated cylinder. where The quasi-two-dimensional character of the electron en- Ϫ eHh cos ␪ 1/2 ergy spectrum in charge-transfer complexes with a layered ␰ϭͩ ͪ ͑ Ϫ ␪ ͒ Py eHx cos /c , structure makes possible an extremely clear manifestation of c the SdH effect, since a considerably greater number of con- then the position of the energy levels of the conduction elec- ␧ duction electrons at the Fermi energy F are involved in the trons is independent of the position of the center of the orbit, formation of this effect than in the case of metals, in which ϭ ␪ ␧ x0 cPy /(eH cos ), and the eigenvalues N of the operator relatively few of the charge carriers are involved in the for- ␧ (Pˆ , P ϪeHx cos ␪/c) are also independent of the integral ⍀ ␧ 1/2 0 x y mation of the effect, a fraction of the order of (h / F) , of the motion of the electrons in the magnetic field, pH where ⍀ϭeH/m*c is their gyration frequency along the ϭ ␪ϩ ␪ Pz cos Py sin . The energy levels have an even simpler closed orbit in the magnetic field, e and m* are the charge ␧ form when 0(px , py) is a quadratic function of the mo- and cyclotron effective mass of a conduction electron, c is menta and all the functions ␧ (p , p ) with nу1 are equal the speed of light, and h is Planck’s constant. n x y to An , i.e., independent of px and py . In that case the energy The substantial dependence of the magnetoresistance of levels of the conduction electrons have the form layered conductors on the orientation of the magnetic field ϱ relative to the layers is inherent only to quasi-two- anp ␧ ͑ ͒ϭ͑ ϩ ͒ ⍀ϩ ͩ H ͪ dimensional conductors, and the position of the maxima of N pH N 1/2 gh ͚ An cos nϭ1 h cos ␪ the function ␳(␪) gives detailed information about the shape of the Fermi surface.10 In contrast to the SdH effect, the ϱ 1 anpH Ϫ ͚ ͑A n͒2␣ sin2ͩ ͪ , ͑3͒ orientational effect is not observed in ordinary metals. 2 n ␪ g h⍀ nϭ1 h cos It is easy to see that not only the smoothly varying ͑with H) part of the magnetoresistance but also the amplitude of where the SdH oscillations also have sharp peaks at certain angles ␪ϭ␪ eH sin2 ␪ c . ␣ϭa2 , We shall assume, solely for the sake of brevity in the hc cos ␪ calculations, that the Fermi surface of a layered conductor ϱ ͑4͒ 2 anp 1/2 consists of only one corrugated cylinder of arbitrary shape. ϭͫ Ϫ ␣ ͩ H ͪͬ g 1 ͚ Ann cos , All of the sections of this cylinder by a plane p"Hϭconst are h⍀ nϭ1 h cos ␪ closed if ␪ differs from ␲/2. The quantum energy levels of and the gyration frequency ⍀ of the charge carriers in the the conduction electrons are determined with the aid of the ¨ magnetic field is the same on all cross sections of the isoen- Schrodinger equation ϭ ergy surface pH const. In the case of a more complicated ␧ dependence of 0 on px and py the energy spectrum is equi- ␧ ˆ Ϫ ␪ ϩ ␪ ͒ distant only at large N (⌬␧ ϭ␧ ϩ Ϫ␧ ϭh⍀). ͑Px ,Py eHx cos /c, Pz eHx sin /c N N 1 N The relation between the current density and the electric ϫ␾ ͑ ͒ ͑ ϩ ͒ N x exp iyPy /h izPz /h field E, ϭ␧ ͑P , P ͒␾ ͑x͒exp͑iyP /hϩizP /h͒ ϭ ˆ ϭ␴ ͑ ͒ N y z N y z ji Tr͕evˆ i f ͖ ikEk 5 ϭ␧ ͑ ͒␾ ͑ ͒ ͑ ϩ ͒ ͑ ͒ N Py , Pz N x exp iyPy /h izPz /h . 2 can easily be found by solving the quantum kinetic equation ˆϭˆ ϩˆ ˆ for the statistical operator f f 0 f 1, where f 0 is the statis- Here we have used the Landau gauge, assuming that the tical operator describing the equilibrium state of the system ˆ NNЈ vector potential A of the magnetic field depends only on the of electrons and has nonzero components f 0 only on the coordinate x, so that the Schro¨dinger equation ͑2͒ contains diagonal, where they are equal to the Fermi distribution ˆ function of the charge carriers f ͓␧ (P , P )͔, and the op- only one differential operator Px and the generalized mo- 0 N y z ˆ mentum projections Py and Pz are good quantum numbers. erator f 1 describes the perturbation of the electron system by The Hamiltonian operator is obtained by replacing the kine- the electric field; vˆ is the electron velocity operator. matic momentum p in formula ͑1͒ for the energy of the In the linear approximation in the weak electric field the charge-carrying quasiparticles by the expression Pˆ ϪeA/c, kinetic equation has the form11–13 and the phase ␣ (p ,p ) has been dropped in Eq. ͑2͒.To n x y i ␾ ϩ ͑␧ Ϫ␧ ͒ NNЈϩ ˆ ͑ ˆ ͒ find the wave functions (x)exp(iyPy /h izPz /h) of the con- N NЈ f 1 WNNЈ f 1 duction electrons, one must specify a concrete form of the h Hamiltonian, and the problem of determining the electron f ͑␧ ͒Ϫ f ͑␧ ͒ ϭ " 0 N 0 NЈ ͑ ͒ energy spectrum in the quasiclassical approximation can be eE vNNЈ ␧ ␧ , 6 ␧ N- NЈ solved for an arbitrary form of the functions n(px , py). ␧ ˆ ˆ If it is assumed that the functions n(px , py) with n where W( f 1) is a linear operator describing the scattering of у ␧ ˆ 1 are independent of px , i.e., that only 0(Px , Py conduction electrons on defects of the crystal and vibrations ϪeHx cos ␪/c) contains the differentiation operator of the crystal lattice. Low Temp. Phys. 27 (99–10), September–October 2001 V. G. Peschansky and R. Atalla 697

͑ ͒ ϭ¯ ␧ ␪ In calculating the oscillatory in the field components in vH(pH , )cos , which we calculate in the quasiclassical ap- ͑ ͒ ⌬␧ Ӷ␩␧ ͑ ͒ the expression for the current density 5 one must take con- proximation ( N F), using the energy spectrum 1 : sideration that the relaxation time of the conduction electrons ϱ exhibits quantum oscillations which arise in the summation 1 T an ¯ ͑ ␧͒ϭϪ ͵ ␧ ͑ ͑ ͒͒ vz pH , dt͚ n py t, pH over the states of the electrons in the incoming term of the T 0 nϭ1 h collision integral Wˆ . Their role is extremely important in the an pH calculation of the asymptotes of the conductivity tensor com- ϫsinͭ ͫ Ϫp ͑t,p ͒tan ␪ͬͮ , ͑9͒ ␴ ␴ ⍀␶ӷ h cos ␪ y H ponents ix and xj in a high magnetic field ( 1). How- ever, their contribution to the amplitude of the quantum os- where Tϭ2␲/⍀, and t is the time of motion of the charge in cillations of the remaining components of the conductivity the magnetic field H, according to the equations tensor is significant only for the small corrections in the pa- ϭ͑ ͒͑ ␪Ϫ ␪͒ ץ ץ rameter ␥ϭ1/⍀␶. In cases where the asymptote of the con- px / t eH/c vy cos vz sin ; ductivity tensor components is nonzero in a high magnetic tϭϪeHv cos ␪/c; ͑10͒ץ/ pץ field and it is unnecessary to keep the small corrections in the y x ϭ ␪ ץ ץ ⍀␶ parameter 1/ , it is sufficient to know only the diagonal pz / t evxH sin /c. matrix elements of the nonequilibrium statistical operator. ␴ For tan ␪ӷ1 the integrand in formula ͑9͒ is rapidly os- The asymptote of the conductivity tensor component zz for ⍀␶ӷ1 has the form cillating with t, and the main contribution to the integral is from small neighborhoods of the points of stationary phase, ␧ ͒ ͑ ͒ ϭ ץ ϱ 2eH f 0͑ N at which, according to Eq. 10 , vx 0. We assume that all ␴ ϭ ͚ dp e2␶¯v2 , ͑7͒ -␧ the cross sections of the Fermi surface have only two stationץ zz 2 H z c͑2␲h͒ Nϭ0 N ary phase points, t1 and t2, the distance between which is where ¯v is the velocity value averaged over P at constant equal to the diameter of the electron orbit along the py axis: z y ϭ Ϫ p , i.e., the average value of v on an electron orbit D p py(t2) py(t1). H z ␪ ␧ϭconst, p ϭconst, and ␶ is the mean free time of the In the case when the angle is substantially different H ␲ charge carriers, which is equal to their inverse collision fre- from /2, all of the quantities appearing in the expression for ␴mon ¯ 2 quency. zz except vz depend weakly on pH , and, dropping terms ␴ In calculating zz we have made use of the solution of of higher order in the quasi-two-dimensionality parameter the kinetic equation in the ␶ approximation, thus giving up than ␩2, we obtain for reliability in relatively unimportant numerical factors of or- f ͑␧͒ ץ 2 der unity which do not affect ones ability to extract informa- ␴monϭϪ ͵ ␧ 0 ͵ ␲ 2␶¯ 2 ͑ ͒ zz d dpH2 m*e vz , 11 ␧ץ tion about the form of the electron energy spectrum from the ͑2␲h͒3 galvanomagnetic characteristics. the following asymptotic expression: In a rather strong magnetic field, when not only ⍀␶ӷ1 but also h⍀/␧ у␩ϭ␩ , the periodic dependence of the ki- F 1 ␴monϭ 2␶ ␪ ␲ 4 2 2͑␪͒ ͑ ͒ netic coefficients on 1/H is rather complicated. However, in zz ae m* cos /2 h ͚ n In , 12 ⍀ ␧ Ӷ␩ the opposite limiting case h / F , this dependence has a harmonic form and can easily be separated off in the expres- where sion for the current density with the aid of the Poisson T ͑␪͒ϭ Ϫ1 ͵ ␧ ͑ ͒ ͑ ͑ ͒ ␪ ͒ ͑ ͒ formula: In T dt n t cos py t an tan /h . 13 0 ␧͒͑ ץ ϱ 2 f 0 The functions I (␪) have a set of zeroes which for ␴ ϭ ʳ ͚ ͵ d␧ ͵ dp 2␲m*e2␶¯v2 n ␧ H z ␪ӷ ⌬ ␪ץ zz 3 ͑2␲h͒ kϭϪϱ tan 1 repeat periodically with period (tan ) ϭ ␲ ͑ ͒ 2 h/naDp . Of course, all of the terms in formula 12 ϫ ͑ ␲ ͑␧ ͒͒ ͑ ͒ exp 2 ikN ,pH . 8 cannot go to zero simultaneously, and the asymptotic behav- ior of ␴mon depends substantially on the character of the The drift velocity ¯v of the conduction electrons along zz H falloff of the functions ␧ (p ,p ) with increasing number n. the magnetic field for (␲/2Ϫ␪)ӷ␩ is weakly dependent on n z y In the tight-binding approximation, when the overlap of the the momentum projection p on the magnetic field direction. H wave functions is small for electrons belonging to different It follows that the expansion of the conductivity tensor com- layers separated by a distance a much greater than the inter- ponents in power series in the quasi-two-dimensionality pa- atomic distance within a layer, A is proportional to the nth rameter ␩ starts with the quadratic terms at least, provided n power of the overlap integral of the wave functions, and ␩ that at least one of the indices of ␴ is z. Then, for deter- n ij is equal in order of magnitude to ␩n. In this case for ␩Ӷ1 mining the resistance to current in the direction transverse to ␥Ӷ ␴mon and 1 the asymptotic expression for zz takes the the layers it is sufficient to know only the conductivity tensor 14 ␴ ␳ form component zz , and only in a few special cases does zz ␴ ͑ ͒ ␴monϭ͑ 2␶ ␪ ␲ 4͒ 2͑␪͒ depend on a larger set of components ij Ref. 14 . zz ae m* cos /2 h I1 For ␩Ӷ1 and ␥ϭ1/⍀␶Ӷ1 the asymptotes of the oscil- ϩ␩2␴ ͑␩2␸ ϩ␥2␸ ͒ ͑ ͒ ͑ ␴osc 0 1 2 , 14 latory in 1/H) part of the conductivity zz and of the ␴mon ␴ part zz which varies monotonically with the magnetic where 0 is the conductivity along the layers in the absence ¯ ␧ field depend substantially on the quantity vz(pH , ) of magnetic field. 698 Low Temp. Phys. 27 (99–10), September–October 2001 V. G. Peschansky and R. Atalla

␸ The functions i are of the order of unity and must be Since the cyclotron effective mass of the conduction elec- ␧ץ ץ ␪ϭ␪ ␪ ϭ ␲ taken into account at those values c at which I1( ) goes trons m* (1/2 )( S/ ) is always nonzero, the condition ␪ϭ␪ ␧ to zero. At c the resistance along the normal to the lay- of stationarity, when S( , pH) is almost constant, can only be ␧ץ ץ ץ ץ ϭϪ ץ ␧ץ ers, rather than saturating in high magnetic fields, increases met for / pH ( S/ pH)( S/ ). Drift of the charge in proportion to H2 in the region of fields satisfying the carriers is absent on the extremal cross sections of the isoen- ϭ ץ ץ ␩Ӷ␥Ӷ condition 1, and saturation of the resistance sets in at ergy surface, i.e., for S/ pH 0, and on the self-crossing higher magnetic fields for ␥р␩. As a result, the angle de- orbit, when m*ϭϱ. For ␩ tan ␪Ӷ1 there are no self- pendence of the resistance to current along the normal to the crossing orbits on the Fermi surface, and ץ ץ layers is expected to exhibit rather sharp peaks. For tan ␪ӷ S/ pH 1 these peaks repeat periodically with period ¯v ͑p , ␧ ͒ϭ cos ␪, ͑20͒ ␧ץ/Sץ z H F ⌬ ␪͒ϭ ␲ ͑ ͒ ͑tan 2 h/aDp . 15 goes to zero only on the extremal cross sections of the Fermi Although it is unclear to what extent the tight-binding surface. After integrating expression ͑16͒ by parts with al- approximation is valid for calculating the electron energy lowance for relation ͑20͒, we obtain spectrum of an extremely complex crystal structure contain- ␧͒ ץ ϱ ing more than a hundred atoms in the unit cell, we shall 2 f 0͑ eHh ␧ ␴oscϭ 2Re͚ ͵ d␧ e2␶ ␧ץ nevertheless assume that the functions n(px , py) fall off, zz 3 ͑2␲h͒ kϭ1 ikc and quite rapidly, with increasing n. ␧ ͒ Although the experimentally observed resistance peaks ikcS͑ , pH ϫ ͵ dp expͩ ͪ are sharp, their height is comparable in order of magnitude H eHh to, and even lower than, the resistance at the minimum of its ϱ 2 dependence on the angle between the normal to the layers an anpH 7,8 ϫ ͩ ͪ ͩ ͪ ͑␪͒ ͑ ͒ ͚ cos In . 21 and the magnetic field direction. This may be because of a ϭ h h cos ␪ ␧ n 1 slow falloff of the functions n(px , py) with increasing n. ␪ ϭϪ␲ ϩ␲ ␪ ␴osc For aDp tan( /2h) /4 k, when I1( ) goes to zero, From here the calculation of zz by the method of sta- ␪ Ϫ all of the In( ) for which (n 1) is not a multiple of four are tionary phase does not present any difficulty. As a result, the substantially nonzero. The low height of the peaks in the oscillatory part of the conductivity transverse to the layers angle dependence of the magnetoresistance attests to a takes the form ␧ slower falloff of the functions n(px , py) with increasing n ϱ eHh 3/2 k␭ 4␲1/2a2 than in the tight-binding approximation, i.e., there are no ␴oscϭ 2␶ͩ ͪ zz ͚ e 2 1/2͉ 2 ץ 2ץ␩ ␩ ϭ kc sinh͑k␭͒ ͑ ␲ ͒3͉ grounds for assuming that the ratio 2 / 1 is much less than k 1 2 h S/ pH h у unity, and to ignore terms with n 2 in the dispersion rela- ϱ ␲ kc͑S ϪS tion of the charge carriers is no longer correct. ϫ Ϫ ͒2 Ϫ max min ͑ ͫ ͚ ͑2n 1 sinͩ ͪ In the expression for the oscillatory in 1/H) part of the nϭ1 4 2eHh conductivity transverse to the layers, kc͑S ϩS ͒ ϱ ϫ ͩ max min ͪ ͑␪͒ ␧͒ cos I Ϫ ץ 2 f 0͑ 2n 1 ␴oscϭϪ 2Re͚ ͵ d␧ 2eHh ␧ץ zz 3 ͑2␲h͒ kϭ1 ϱ ␲ Ϫ ͒ kc͑Smax Smin ϩ ͚ ͑2n͒2cosͩ Ϫ ͪ ϫ ͵ ␲ 2␶¯ 2 ͑ ␲ ͑␧ ͒͒ ͑ ͒ nϭ1 4 2eHh dpH2 m*e vz exp 2 ikN ,pH , 16 ϩ ͒ kc͑Smax Smin there is substantial dependence on pH not only in the factor ϫ ͩ ͪ ͑␪͒ͬ ͑ ͒ sin I2n , 22 ϱ 2eHh an anp ¯ ͑ ␧͒ϭϪ ͩ H ͪ ͑␪͒ ͑ ͒ ␭ϭ ␲2⌰ ⍀ ⌰ vz pH , ͚ sin In , 17 where 2 /h , and is the temperature in energy ϭ h h cos ␪ n 1 units. ͑ ប⍀Ӷ␩␧ ␧ but also in the oscillatory for F) factor If the function n(px , py) falls off quite rapidly with ␲ ␧ ␩ Ӷ␩ exp(2 ikN( , pH)). increasing index n, i.e., nϩ1 n , then the substantial de- ␴osc ␴osc ␪ϭ␪ ␴mon In calculating zz in the quasiclassical approximation crease of zz sets in at the same angles c as for zz , one should use the area quantization rule4 and the amplitude of the SdH oscillations of the magnetore- ␳osc ␴osc ␴mon 2 sistance, zz , being proportional to zz /( zz ) , increases 1 eH 1 ␪ϭ␪ S͑␧, p ͒ϭ S ͑␧, p ͒ϭ2␲h ͩ Nϩ ͪ , ͑18͒ rapidly at c . As a result, sharp peaks appear in the de- H cos ␪ 0 H c 2 ␳osc ␪ pendence of zz on the angle between the vectors H and n. ␧ ϭ͛ If ␩ is not much smaller than ␩ , i.e., if I (␪) and I (␪) are where S0( , pH) pydpx is the projection on the (pxpy) 2 1 2 1 ␧ quantities of approximately the same order, then the peaks in plane of the section of the isoenergy surface S( , pH) by the osc ␳osc plane p ϭconst. The main contribution to ␴ comes from the angle dependence of zz will be shifted from the peaks H zz ␳mon integration of small neighborhoods near the points of station- of zz . From the value of this shift one can estimate the ary phase, which are found from the condition quasi-two-dimensionality parameter of the electron energy spectrum. S c ␴osc ␴monץ Sץ ͑␧ ͒ϭͩ ␧ϩ ͪ ϭ ͑ ͒ The formulas for zz and zz given above are valid for dpH ␲ 0. 19 ␪ ␥Х␥ ␪Ӷ ץ ␧ dץ dN , pH pH 2 heH not too large values of tan , when 0 /cos 1, and in Low Temp. Phys. 27 (99–10), September–October 2001 V. G. Peschansky and R. Atalla 699 the oscillation amplitude one may neglect the Dingle factor stantially alter the asymptotic behavior at small ␩ not only of Ϫ␥ ␳mon exp( ), which takes into account the broadening of the zz but also of the amplitude of the SdH oscillations of the ␳osc quantum energy levels due to scattering of the charge magnetoresistance, zz . carriers.15 Here ␥ ϭ1/⍀ ␶, and ⍀ is the gyration frequency 0 0 0 * of a conduction electron in a magnetic field directed along E-mail: [email protected] the normal to the layers. When ␪ approaches ␲/2 the orbits become highly elon- 1 L. V. Schubnikov and W. J. de Haas, Leiden Commun. 19, 207f ͑1930͒. ␥ ␪у 2 ͑ ͒ gated, and for 0 tan 1 the electron does not have time to W. J. de Haas, J. W. Blom, and L. V. Schubnikov, Physica 2,907 1930 . 3 complete a full revolution along the orbit in the magnetic L. Onsager, Philos. Mag. 43, 1006 ͑1952͒. 4 I. M. Lifshits and A. M. Kosevich, Zh. E´ ksp. Teor. Fiz. 29,730͑1955͒ field. In this region of angles the amplitude of the magne- ͓Sov. Phys. JETP 2, 636 ͑1956͔͒. toresistance oscillations decays exponentially with increasing 5 W. J. de Haas and P. M. van Alphen, Leiden Commun. 19, 208d ͑1930͒. ␪, while ␳mon , on the contrary, increases sharply when ␪ 6 L. D. Landau, Proc. R. Soc. London, Ser. A 170, 341 ͑1939͒. zz 7 approaches ␲/2 for (␲/2Ϫ␪)Ӷ␩.16,17 This is because a con- M. V. Kartsovnik, V. N. Laukhin, V. N. Nizhankovskiy, and A. A. Ignat’ev, JETP Lett. 47,363͑1988͒. striction along the px axis appears on the electron orbit with 8 M. V. Kartsovnik, P. A. Kononovich, V. N. Laukhin, and I. F. Shchegolev, increasing ␪, the value of which, ⌬p, tends toward zero JETP Lett. 48,541͑1988͒. when ␩ tan ␪ is of the order of unity. The time for traversing 9 I. D. Parker, D. D. Pigram, R. H. Friend, M. Kurmo, and P. Day, Synth. ⌬ Met. 27, A387 ͑1988͒. this constriction is large and diverges logarithmically as p 10 ␥ ␪ V. G. Peschansky, J. A. Roldan Lopez, and Toji Gnado, J. Phys. I 1, 1469 approaches zero. As a result, increases with increasing ͑1991͒. more rapidly than tan ␪, and the resistance transverse to the 11 I. M. Lifshits, Zh. E´ ksp. Teor. Fiz. 32, 1509 ͑1957͓͒Sov. Phys. JETP 5, layers initially falls off with increasing ␪. The drift velocity 1227 ͑1957͔͒. 12 ͑ ͒ ¯ E. Adams and T. Holstein, J. Phys. Chem. Solids 10,254 1959 . of the charge carriers along the normal to the layers, vz ,is 13 A. M. Kosevich and V. V. Andreev, Zh. E´ ksp. Teor. Fiz. 38,882͑1960͒ ␪ ␳ ͓ ͑ ͔͒ proportional to cos , and the resistance zz , after passing Sov. Phys. JETP 11, 637 1960 . through a minimum in the region of angles where cos ␪ is of 14 V. G. Peschanskiy, Fiz. Nizk. Temp. 23,47͑1997͓͒Low Temp. Phys. 23, ͑ ͔͒ the order of ␩, again increases, reaching its maximum value 35 1997 . 15 Yu. A. Bychkov, E´ ksp. Teor. Fiz. 39, 1401 ͑1960͓͒Sov. Phys. JETP 12, at ␪ϭ␲/2. 971 ͑1961͔͒. ␳ 16 Thus the resistivity zz of a layered conductor along the M. V. Kartsovnik and V. G. Peschansky, J. Low Temp. Phys. 117,1717 ‘‘hard’’ direction of current flow, i.e., along the normal to the ͑1999͒. 17 ͑ ͒ layers, is extremely sensitive to the orientation of the mag- V. G. Peschansky and M. V. Kartsovnik, Phys. Rev. B 60, 11207 1999 . netic field, and at certain values of the angle ␪ it can sub- Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Magnetization and magnetostriction oscillations in a superconducting 2H-NbSe2 single crystal V. Eremenko, V. Sirenko,* and Yu. Shabakayeva

B. Verkin Institute for Low Temperature Physics and Engineering of the National Academy of Sciences of Ukraine, 47 Lenin Ave., Kharkov 61103, Ukraine R. Schleser

Grenoble High Magnetic Field Laboratory, MPI-FKF and CNRS, BP-166, F-38042, Grenoble Cedex 09, France P. L. Gammel

Bell Laboratories, Lucent Technologies, Murray Hill, New Jersey 07974, USA ͑Submitted May 29, 2001͒ Fiz. Nizk. Temp. 27, 952–955 ͑September–October 2001͒ The results of a comparative study of the oscillating magnetostriction and absolute magnetization components in the single crystal of the layered superconducting compound 2H-NbSe2 are presented. The measurements were made in the temperature range 1.5–8 K in a magnetic field of 0–20 T with in-plane orientation ͑normal to the hexagonal axis͒. The mechanical stress derivative of the extremal cross-sectional area of the Fermi surface is derived from a thermodynamic analysis. The values obtained are considerably higher, by an order of magnitude, than those for other metals, which explains the observed enhancement of the magnetostriction oscillations in comparison with the magnetization oscillations. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1401177͔

1. INTRODUCTION nounced dependence F(␪) in the angle range 0Ͻ␪Ͻ20° and a weak dependence F(␪)at␪Ͼ60°.6 Band calculations5 L. V. Shubnikov was the first to observe oscillations of demonstrate that the measured angular dependence of the the resistivity.1 It was this observation that led to the discov- ery of the de Haas–van Alphen ͑dHvA͒ oscillations of the oscillations is well described in terms of a ‘‘small, flat hole Fermi surface around the ⌫ point.’’ The above-mentioned magnetic properties. The layered dischalcogenide 2H-NbSe2, with a transition to the charge-density-wave ͑CDW͒ state at investigation complemented the results of previous band ϭ calculations9–11 of the Fermi surface in 2H-NbSe . Existing temperature TCDW 32 K, is the first superconducting com- 2 ͑ pound the superconducting transition temperature Tc data about the shape and peculiarities of the Fermi surface in ϭ7.2 K͒ in which Landau oscillations of the electron spec- 2H-NbSe2 suggest its high sensitivity to pressure and stress, trum have been observed.2 That was the first observation of which has been proved by the measured pressure depen- the de Haas-van Alphen and magnetothermal oscillations not dences of the superconducting parameters of this compound only in the normal state but also in a mixed state, the so- ͑Refs. 12–14 and references therein͒. The observed magne- Ͻ 15 called Shubnikov phase (H Hc2). The measurements were tostriction oscillations in superconducting single crystals of made at temperatures much lower than T for different ori- c 2H-NbSe2 in comparison with dHvA oscillations can be used entations of the external magnetic field H with respect to the to study the influence of mechanical stresses on the band crystallographic axes of the single crystal under study. In that structure of this material according to the thermodynamic 2 3–7 study and in subsequent investigations of superconduct- approach developed in16–18. ing 2H-NbSe2 the measurements of the magnetization oscil- We should mention that in19 the effect of deformation on lations were made at a temperature far below1Kinorder to dHvA oscillations is considered. The related change of oscil- eliminate the electron scattering as above the Dingle tem- Ϫ Ϫ lation frequency is estimated. Deformations of 10 2 –10 3 perature ͑T ϭ2.5 K, ͑Ref. 5͒, T ϭ1K ͑Ref. 6͒ for D D are taken into account. Here, we analyze the amplitude ratio 2H-NbSe ͒ it reduces drastically the amplitude of the oscil- 2 for the magnetization and magnetostriction, the amplitude of lations, and their registration by existing techniques becomes Ϫ5 impossible. Measurements in magnetic fields below the up- the latter being of the order of 10 . An analysis of the frequencies will be presented elsewhere. per critical field Hc2 have shown that in the mixed state the oscillations are suppressed by the scattering of normal ͑quan- In this work a comparative analysis of measurements of ͑␭ tized͒ electrons on Abrikosov vortices, observed in other su- the oscillatory magnetostriction OSC and absolute magneti- perconductors also ͑see the review8 and the references zation Mϭm/V ͑m is the magnetic moment of the sample therein͒. The measured angular dependences of the oscilla- and V is its volume͒ is performed. The same single crystal of tion frequency F(␪), where ␪ is the angle between the hex- 2H-NbSe2 was studied by capacitive dilatometer and capaci- Ͻ 20 agonal axis and field direction, at H Hc2 revealed a pro- tive cantilever torquemeter techniques.

1063-777X/2001/27(9–10)/4/$20.00700 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 Eremenko et al. 701

FIG. 1. Capacitive torquemeter for magnetization measurements. The cali- bration coil is positioned in the sample area ͑not shown here͒.

2. EXPERIMENTAL TECHNIQUES The absolute magnetization measurements were per- formed on single crystals of the layered compound ͑ 4 ͒ 2H-NbSe2 space group P6/mmm or D6h . The high anisot- ropy of the crystal structure ͑aϭ3.45 Å, cϭ12.54 Å͒ is ac- companied by anisotropy of the electronic properties, in par- ϭ ticular, of the superconducting parameters: Hc2(T 4.2 K) Ϸ7 T for ␪ϭ90° and 2.3 T for ␪ϭ0° ͑there is some discrep- ancy in the published data͒.2,15,21,22 The mass of the sample ␭ ͑ ͒ ͑ ͒ ϭ Ϯ FIG. 2. The typical magnetic field dependences of rev a and M rev b for was m (127 0.05) mg. The sharp superconducting transi- 2H-NbSe single crystals. ϭ 2 tion was observed at temperature Tc 7.23 K. The detailed information on the sample preparation technique and charac- terization is presented in Ref. 15. In that paper the first ob- reason the field dependence of M was measured at ␪ϭ(77 servation of magnetostriction oscillations in the mixed state Ϯ1)°. The measurements revealed equal frequencies of the of any superconductor was presented. The measurement of magnetization and magnetostriction oscillations, F ϭF , (c,a) M MS magnetostriction ␭ was performed on a single crystal of which agrees with the weak angular dependence F(␪)at ͑ ͒ 2H-NbSe2 in the crystallographic direction 001 , i.e., the c 60°Ͻ␪Ͻ90°, measured earlier. In order to find the absolute axis, in a magnetic field parallel to the a axis. The oscillatory values of M, calibrations of the measuring capacitance were ␭(c,a) ␭(c,a)ϭ␭(c,a)Ϫ␭(c,a) component of was defined as OSC rev mon , performed in each experiment using a calibrating coil, which ␭(c,a)ϭ ␭ ϩ␭ ␭(c,a) where rev ( up down)/2 is the reversible part of ; was positioned in the region of the sample ͑Fig. 1͒. The ␭(c,a) ␭(c,a) up is measured during magnetization of the sample; volume V of the sample was determined using the density ␭(c,a) ␭(c,a) 3 3 5 down is measured during demagnetization of the value ␳ϭ6.3 10 kG/m , calculated from the known atomic ␭(c,a) ␭(c,a) • sample; mon is the monotonic part of rev . distribution in 2H-NbSe2. The oscillatory component of M In the present work the absolute magnetization measure- was defined in the same way as the oscillatory component of ments were performed in a temperature range 1.5–4.2 K in a ϭ ϩ Ϫ ¯ the magnetostriction: M osc 1/2(M up M down) M, where magnetic field up to 20 T using a capacitive cantilever ϭ ϩ M rev 1/2(M up M down) is the reversible component of M, torquemeter ͑see Fig. 1͒. The field sweep rate was 0.18 M is the absolute magnetization of the sample measured in 22 up T/min. The torque magnetization measuring technique is increasing magnetic field after zero field cooling ͑ZFC͒; based on the fact that in a magnetic field H a magnetized M down is the absolute magnetization of the sample measured sample experiences a torque T proportional to the compo- in increasing magnetic field; M¯ is the monotonic component nent of the sample’s magnetic moment perpendicular to H: of M. In the oscillatory regime it was determined by averag- V ing M osc . Tϭ MϫH, ͑1͒ ␮ Analysis of M osc was performed after subtracting the 0 monotonic component according to Ref. 6. ␮ where 0 is the vacuum magnetic permeability. In a sample with an anisotropic Fermi surface the mag- 3. EXPERIMENTAL RESULTS AND DISCUSSION netization component MЌ perpendicular to the field direction 23 is proportional to the parallel component M ʈ : Figure 2 presents a typical magnetic-field curve of M rev ␭ -F for a 2H-NbSe2 single crystal, in comparison with rev meaץ 1 ϭϪ ϭ ͑ ͒ ␪ 2 sured on the same crystal. It is clearly seen that in a region ofץ MЌ kM, k F the peak effect15 both M and ␭ depend nonlinearly on H. where F is the free energy. This means that a simple summation of the dependences The torque T results in deflection of a Cu-Be spring measured in increased and decreased fields does not com- joined with the sample. The change of the distance between pletely eliminate manifestations of the peak effect on the ␭ the spring and adjacent metallic plate is measured by a ca- M rev(H) and rev(H) curves. The oscillatory behavior of the ␪ϭ pacitance technique. At 90° the measured signal from the measured dependences starts near Hc2 . Unfortunately, be- torquemeter disappears, as k becomes equal to zero. For this cause of the available experimental conditions the magneto- 702 Low Temp. Phys. 27 (99–10), September–October 2001 Eremenko et al.

from a review18 for common metals. ␴ץ ץ It is seen that the values of (lnSm)/ ␣ obtained in the present work for 2H-NbSe2 are considerably, by an order of magnitude, higher than those obtained previously for the other metals. This is why the magnetostriction oscillations are enhanced in comparison with the magnetization oscilla- tions.

CONCLUSIONS A comparative analysis of the reversible absolute mag- netization and magnetostriction as functions of the inverse magnetic field, measured on a superconducting 2H-NbSe2 single crystal, have shown that: 1͒ the oscillation periods ␭ M osc and osc coincide, which means that both effects are attributed to Landau quantum oscillations; 2͒ the magneto- striction oscillations are more pronounced than the magneti- zation oscillations: the former appear at lower fields, and their amplitude is higher ͑with respect to the monotonic com- ponent͒;3͒ a thermodynamic analysis shows that the values ␴ץ ץ of (lnSm)/ ␣ for 2H-NbSe2 are much higher, by an order ␴␣ for commonץ/( lnS)ץ of magnitude, than the values of ␭ ͑ ͒ ͑ ͒ m FIG. 3. The inverse field dependences of osc a and M osc b for T metals. This is due to the peculiar geometry of Fermi surface ϭ1.5 K. The inset shows the results of a fast Fourier transform analysis of in the highly anisotropic compound 2H-NbSe and explains the measurements, which was used for determination of F M and F MS . 2 the effect of magnetostriction ‘‘enhancement.’’ The authors are indebted to P. Wyder for support of the striction measurements were restricted by an upper value of work, to T. J. B. Janseen for some stimulating discussions, the applied field Hϭ14 T, and comparison with the magne- and to N. Makedonskaya for assistance in preparation of the tization measurements is possible only for this range of H. manuscript. The magnetization oscillations are less pronounced than those of the magnetostriction. They appear at higher fields, *E-mail: [email protected] and their amplitude AM is lower than the amplitude AMS of the magnetostriction oscillations. Figure 3 presents the in- ␭ ϭ verse field dependences of M osc and OSC for T 1.5 K, which were obtained by subtracting their monotonic compo- 1 L. V. Schubnikov and W. J. Haas, Proc. Netherlands R. Acad. Sci. 33,163 nents from M and ␭ , respectively. The insets show the ͑1930͒. rev rev 2 J. E. Graebner and M. Robbins, Phys. Rev. Lett. 36, 422 ͑1976͒. results of a fast Fourier transform analysis of the measure- 3 Y. Onuki, I. Umehara, N. Ebihara, N. Nagai, and K. Takita, J. Phys. Soc. ments, which was used for determination of FM and FMS .It Jpn. 61, 692 ͑1992͒. is seen that these frequencies coincide. The slight phase shift 4 Y. Onuki, I. Umehara, T. Ebihara, A. K. Albessars, K. Satoh, K. Takita, H. ͑ ͒ may be explained by the different geometry of the experi- Aoki, S. Uji, and T. Shimuzu, Physica B 150, 186 1993 . 5 R. Corcorant, P. Meeson, Y. Onuki, P. A. Probst, M. Springford, K. Toki- ment, explained in the previous Section. The derived values tat, H. Harimot, G. V. Guo, and B. L. Gyoffy, J. Phys.: Condens. Matter 6, ␴ץ ץ of AM and AMS were used for calculations of (lnSm)/ ␣ , 4479 ͑1994͒. 6 S. Rettenberger, E. Steep, F. Meyer, A. G. U. Jansen, W. Jose, P. Wyder, where Sm is the extremal cross-sectional area of the Fermi W. Biberacher, E. Bucher, and C. S. Oglisby, Physica B 211,244͑1995͒. surface, and ␴␣ is the mechanical stress arising along the H 7 E. Steep, S. Rettenberger, F. Meyer, A. G. U. Jansen, W. Jose, W. Biber- direction. For that we used the expression acher, E. Bucher, and C. S. Oglisby, Physica B 204, 162 ͑1995͒. 8 T. J. B. Janssen, C. Haworth, S. M. Hayden, P. Messon, and M. Spring- ͒ ln S͑ץ ␭ ϭ m ͑ ͒ ford, Phys. Rev. B 57, 11698 ͑1998͒. , 3 9 ͑ ͒ . ␴␣ L. F. Matheises, Phys. Rev. B 8, 3719 1973ץ M H • 10 C. Y. Fong and M. L. Cohent, Phys. Rev. Lett. 32,720͑1974͒. which was obtained from a thermodynamic analysis of the 11 F. G. Wexler and A. Wooly, J. Phys. C 9, 1185 ͑1976͒. oscillating part of the thermodynamic potential.16–18 The re- 12 M. A. Obolenskii, Kh. B. Chashka, V. I. Belestkii, and V. M. Gvozdikov, ͑ ͓͒ sults are presented in Table I in comparison with the data Fiz. Nizk. Temp. 15, 984, 1989 Sov.J.LowTemp.Phys.15,544 ͑1989͔͒. 13 Kazuhiko Yamaya and Takashi Samboni, J. Phys. Soc. Jpn. 32,1150 ͑1972͒. ␴ 14ץ ץ TABLE I. Comparison of (ln Sm)/ ␣ values for 2H-NbSe2 and other V. G. Kogan, L. N. Bulaevskii, P. Miranovic, and L. Dobrosavljevic- metals. Grujic, Phys. Rev. B 51, 15344 ͑1995͒. 15 V. V. Eremenko, V. A. Sirenko, R. Schleser, and P. L. Gammel, Fiz. Nizk. ͔͒ ͑ ͓͒ ͑ ͒ ␴ ͑ Ϫ11 2ץ ץ Metal (ln Sm)/ ␣ in 10 m /N Temp. 27, 412 2001 Low Temp. Phys. 27, 305 2001 . 16 I. M. Livshits and A. M. Kosevich, Izv. Acad. Nauk 19,395͑1955͒. Ϯ ͓ ͔ 17 2H-NbSe2 148.5 11 present work B. S. Chandrasekhar, J. H. Condon, E. Fawcett, and W. M. Becker, Phys. Cu 7.5Ϯ2 ͓Ref. 18͔ Rev. Lett. 17, 954 ͑1966͒. Ag 24Ϯ1 ͓Ref. 18͔ 18 B. S. Chandrasekhar and E. Fawcett, Adv. Phys. 20, 775 ͑1971͒. Au 8.3Ϯ1 ͓Ref. 18͔ 19 A. M. Kosevich, Zh. E´ ksp. Teor. Fiz. 35, 249 ͑1959͓͒Sov. Phys. JETP 35, 171 ͑1959͔͒. Low Temp. Phys. 27 (99–10), September–October 2001 Eremenko et al. 703

20 P. Christ, W. Biberacher, H. Mueller, and K. Andres, Solid State Commun. 23 D. Shoenberg, Magnetic Oscillations in Metals, Cambridge Univer- 91,451͑1994͒. sity ͑1984͒. 21 P. de Trey, Suso Gydax, and J. P. Jan, J. Low Temp. Phys. 11, 421 ͑1973͒. 22 R. C. Morris, R. V. Coleman, and R. Bhandari, Phys. Rev. B 5, This article was published in English in the original Russian journal. Repro- 895 ͑1972͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Shubnikov–de Haas oscillations in layered conductors with stacking faults V. M. Gvozdikov*

Department of Physics, Kharkov National University, 4 Svobody Sq., 61077 Kharkov, Ukraine; Grenoble High Magnetic Field Laboratory MPI-FRF and CNRS, BP 166 F-38042, Grenoble Cedex 09, France ͑Submitted July 9, 2001͒ Fiz. Nizk. Temp. 27, 956–962 ͑September–October 2001͒ The Shubnikov–de Haas ͑SdH͒ oscillations of the in-plane conductivity of layered 2D electron gas are calculated. It is shown that layer stacking faults, magnetoimpurity bound states, and electron scattering modulate the SdH oscillations via the specific factors which bear the structural information. At zero temperature the 2D SdH oscillations are strongly nonsinusoidal in shape and related by a simple equation to the derivative of the de Haas–van Alphen magnetization oscillations with respect to the magnetic field. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1401178͔

INTRODUCTION which is a Fourier transform of the one-dimensional density of states ͑DOS͒ N(␧) related to electron hopping across the The discovery of high-Tc superconductivity in layered layers ͑p is the number of the harmonic͒. ␧ ϰ␦ ␧Ϫ␧ cuprates and the unconventional electronic properties of lay- The delta peak in the N( ) ( 0) at some energy ␧ ered organic superconductors make structural studies of 0 makes I p an oscillatory function of the inverse field, these layered conductors of great interest. In particular, much of recent experimental effort has been devoted to de Haas– ␲ ␧ van Alphen ͑dHvA͒ and Shubnikov–de Haas ͑SdH͒ studies 2 p 0 I ϳcosͩ ͪ , in organic superconductors.1–7 Both the dHvA and SdH mea- p ប⍀ surements in these quasi-two-dimensional ͑Q2D͒ materials displayed numerous deviations from the conventional three- 8 ␧ dimensional ͑3D͒ theory of Lifshitz and Kosevich ͑LK͒, with the frequency determined by 0 . Another modulation many of which still remain not understood. For example, a factors appear in the case of the magnetic field parallel to the 13 sawtooth profile of the dHvA signal was predicted at zero layers and under the conditions of coherent magnetic 14 temperature Tϭ0 for a purely 2D electron gas in the canoni- breakdown. cal ensemble ͑i.e., for fixed number of electrons in the Impurities play a significant role. In the theory of the sample͒,9 and an inverse sawtooth was found within the dHvA oscillations they smear the Landau levels and thereby 10 grand canonical ensemble ͑i.e., for fixed chemical potential decrease the harmonic amplitudes via the Dingle factor, or in the sample͒.10 In the first case the dHvA oscillations are they may create bound electronic states which manifest 15 due to the oscillations of the chemical potential ␮. The recent themselves through beats of the dHvA oscillations, which 16 analysis of this matter given in Ref. 11 shows that it is only are more prominent in the 2D case. for low electron concentrations, when it is possible to realize Contrary to the dHvA effect the SdH oscillations appear experimentally a regime ប⍀Ϸ␮ ͑⍀ϭeH/mc is the cyclo- as a result of the electron lifetime oscillations due to the 17 tron frequency͒, that oscillations of ␮ with inverse magnetic scattering on impurities in an external magnetic field. field 1/H are important. In the case ␮ӷប⍀ the oscillations Theoretically it is a much more difficult problem than the of the chemical potential are small, and one can use the study of the thermodynamic oscillations and, consequently, grand canonical ensemble as it holds in the standard 3D LK the number of publications on the SdH effect in layered con- theory. ductors is much less than on the dHvA effect in Q2D sys- Real layered conductors are far from being an ideal 2D tems. The SdH effect in the case of a 2D electron gas was electron gas. As a rule they consist of a stack of 2D conduct- considered in a few works reviewed in Ref. 18. The anisot- ing planes with an anisotropic 2D Fermi surface within them ropy in a 2D model was considered numerically in Ref. 19 in and may contain impurities as well as some amount of stack- the context of the SdH studies of the organic superconduct- ing faults appearing as a result of the intercalation. The im- ors. Longitudinal magnetoresistance quantum oscillations pact of the stacking faults on the dHvA oscillations in a due to the coherent hopping across the layers has been nu- layered conductors was studied by the author in Ref. 12. It merically studied in Refs. 20 and 21 for the incoherent case, was shown that amplitudes of the dHvA harmonics are i.e., when ␶Ͼប/t ͑␶ is the electron lifetime, and t stands for modulated by the layer factor the hopping integral between the neighboring layers͒. In this paper we calculate ␶ in a layered 2D electron gas containing a small amount of stacking faults in an external quantizing magnetic field perpendicular to the layers and ϱ 2␲p␧ ϭ ͵ ͑␧͒ ͩ ͪ ␧ ͑ ͒ then apply the results of these calculations to the theoretical I p N exp i d , 1 Ϫϱ ប⍀ consideration of the SdH oscillations in this system.

1063-777X/2001/27(9–10)/5/$20.00704 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Gvozdikov 705

THE LIFETIME CALCULATION 4␲tp ϭ ͩ ͪ ͑ ͒ I p J0 ប⍀ , 7 A theory for the single-electron lifetime ␶(E)ina2D

electron gas due to the impurity scattering in an external where J0(x) is the Bessel function and t is the interlayer quantizing magnetic field was developed by Ando and hopping integral. 22 Uemura. This theory provided a basis for further studies of In case when there is a small concentration cӶ1 of layer 18 different properties of a 2D electron gas in magnetic fields. stacking faults, the factor I p was calculated in Ref. 12 to be Maniv and Vagner23 have generalized the approach of Ando 4␲tp and Uemura to the case of a layered electron gas. These ϭ͑ Ϫ ͒ ͩ ͪ ϩ ␦ ͑ ͒ I p 1 c J0 c I p , 8 authors have shown that in layered conductors the interplane ប⍀ tunneling of electrons can significantly reduce the 2D scat- where tering rate due to the possibility of escaping the impurity by ␲ ␲ hopping to the neighboring conducting planes. They calcu- 2 p 4 t0 ␦I ϭexpͩ Ϫ ͱt2Ϫt2 ͪ cosͩ pͪ . ͑9͒ lated the magnetic field dependence of ␶ assuming the kinetic p ប⍀ 0 ប⍀ ␧ ϭ 2 energy across the layers (kZ) kZ/2m to be limited within Because of the stacking fault, the local hopping integral р␧ р⌬ the band 0 (kZ) Z . Here we generalize their approach t0 relevant to the electron transition from the ‘‘defect layer’’ to the case when the interlayer hopping energy is an arbitrary to the nearest-neighbor layers is less than t. Physically this ␧ quantity distributed within the DOS N( ). Then the equation kind of stacking faults may appear as a result of the interca- ␴͑␻͒ 23 for the single-electron self-energy yields: lation of the layered conductor. ϱ ϱ N͑␧͒ ␴͑␻͒ϭ ϩ ͵ ␧ ͑ ͒ b0 c0 ͚ d ␧Ϫ␥ ␻͒ϩ␴ ␻͒ , 2 nϭ0 Ϫϱ n͑ ͑ SHUBNIKOV-DE HAAS OSCILLATIONS where Consider now the SdH oscillations in a stack of 2D iso- ␥ ͑␻͒ϭ␻Ϫប⍀͑ ϩ ͒ϩ␮ ͑ ͒ n n 1/2 , 3 tropic conductors with impurities in a perpendicular mag- 2 netic field. We do not assume periodicity in the layer stack- U0ni b ϭU n , c ϭ , ͑4͒ ing. The diagonal elements of the in-plane conductivity 0 0 i 0 ␲2 2 18 4 lH tensor are therefore given by ץ U is the impurity potential amplitude V(r)ϭU ␦(rϪr ), n 2 ϱ 0 0 i i Nee f is the impurity concentration, and l ϭ(បc/eH)1/2 is the ␴␣␣ϭϪ ͵ dEͩ ͪ ␶͑E͒, ͑10͒ Eץ H m magnetic length. 0 ͑ ͒ Separating the real and imaginary parts in Eq. 2 for the where Ne is the 2D density of conduction electrons within self-energy ␴(␻)ϭRe ␴(␻)ϩi Im ␴(␻), we have the layer and f (E) is the Fermi function. ϱ The energy-dependent lifetime ␶(E)ϭប/Im␴(E) can be ϱ ␴͑␻͒ϭ ͵ ␧ ͑␧͒ found from Eq. ͑6͒ to a first approximation in the parameter Im c0 ͚ d N nϭ0 Ϫϱ a/ប⍀Ӷ1, which yields ϱ Im ␴͑␻͒ ប 2␲pa ϫ . ͑5͒ ␶͑ ͒Ϸ ͭ Ϫ ͚ ͑Ϫ ͒p ͩ Ϫ ͪ ␥ ␻͒Ϫ ␴ ␻͒Ϫ␧ 2ϩ 2 ␴ ␻͒ E 1 2 1 exp ͓ n͑ Re ͑ ͔ Im ͑ a pϭ1 ប⍀ Since we are interested in quantum oscillations, which E are relevant to the long-lifetime regime, ប/␶(␻)ϭIm ␴(␻) ϫ ͫ ␲ͩ ϩ␦ͪ ͬͮ ͑ ͒ I p cos 2 p . 11 Ӷប⍀ ␴ ␻ Ӎ Ӷ␮ ប⍀ , we can neglect the quantity Re ( ) U0ni as a small correction to the chemical potential ␮. Substituting Eq. ͑11͒ into Eq. ͑10͒ and completing the stan- 10,17 Then, applying a standard summation procedure to dard integration, we have Eq. ͑5͒, we can recast it in the form ␴ ϭ␴ ϩ␴ ͑ ͒ ␣␣ 0 ˜ , 12 ϱ ␲ ␴͑␻͒ 2 Im ␴ ϭ 2␶ Im ␴͑␻͒ϭaͭ 1ϩ2Re͚ ͑Ϫ1͒p expͫϪ ͬ where 0 Nee 0 /m is the smooth part of the conductivity, pϭ1 ប⍀ ␶ ϭប ϭ ␲2ប3 2 with 0 /a 4 /mU0ni . ␴ ␻ The oscillating part of the conductivity ˜ is given by ϫ ͫ ␲ͩ ϩ␦ͪ ͬͮ ͑ ͒ ϱ I p cos 2 ⍀ p , 6 2␲p 2␲␮ ␴ϭϪ ␴ ͑Ϫ ͒p ͩ Ϫ ͪ T ͩ ͪ ˜ 2 0 ͚ 1 exp R pI p cos p , ϭ ␶ ⍀ ប⍀ where aϭn E /2␲ with E ϭmU2/2ប2 being the resonance p 1 0 i 0 0 0 ͑13͒ energy for the ␦-potential scattering, and ␦ϭ⌬͓1/H͔/ប⍀ is a small phase shift 0Ӷ⌬͓1/H͔Ͻ1/2 arising because the ratio where ␮/ប⍀ is not exactly an integer: ␮/ប⍀ϭNϩ⌬͓1/H͔ ͑N Z p stands for an integer͒. The factor I here is given by Eq. ͑1͒. RTϭ ͑14͒ p p sinh Z This factor was studied in detail in Ref. 12, which is devoted p ϭ ␲2 ប⍀ to the dHvA effect in layered conductors. For a periodic is the standard temperature factor, with Z p 2 Tp/ .We layered system with nearest-layer hopping we have N(␧) also have neglected the small phase shift ␦, since energies of ϭ1/␲(4t2Ϫ␧2)1/2 and the order of EӍ␮ are important in the integral of Eq. ͑10͒. 706 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Gvozdikov

ϭ Considering at first the case of a solitary layer (I p 1) Tϭ and zero temperature (R p 1), one can perform an elemen- tary summation in Eq. ͑14͒ to obtain sinh ␯ ˜␴ϭ␴ 1Ϫ , ͑15͒ 0ͫ 2␲␮ ͬ ␯ϩ cosh cosͩ ប⍀ ͪ ␯ϭ ␲ ⍀␶ where 2 / 0 . On the other hand, the oscillating part of the magnetiza- tion of a 2D electron gas at zero temperature can be written as a sum10 ϱ ͑Ϫ1͒p 2␲p 2␲␮ ˜ ϭ ͩ Ϫ ͪ ͩ ͪ ͑ ͒ M M 0 ͚ exp ␶ ⍀ sin ប⍀ p , 16 pϭ1 p 0 which can be completed to yield 2␲␮ sinͩ ប⍀ ͪ ˜ ϭ ͑ ͒ M M 0 arctanͫ ͬ . 17 2␲␮ ϩ ␯ cosͩ ប⍀ ͪ e

Here M 0 is independent of the magnetic field H. Comparing Eqs. ͑15͒ and ͑17͒ we arrive at the relation between ˜␴ and M˜ for a 2D electron gas with ␦-potential impurities at zero temperature:

͑͒ ͑ ˜M ץ ␴˜ Ϸ 2 ͩ ͪ ͑ ͒ FIG. 1. Field dependence of the conductivity given by Eq. 15 in conven- AH , 18 ϭប⍀͒ ␮ϭ ␯ϭ ␯ϭ ␲ប ␶ .(tional units B for 1000, 0.1, 0.151, 0.2 ( 2 / 0 ץ ␴ 0 H M 0 where Aϭeប/␲mc␮. Since AHӍប⍀/␲␮Ӷ1, we see that ␴ ␴ relative amplitude of the SdH oscillation ˜ / 0 is much less than the corresponding relative oscillations of the magneti- ϱ 2␲␨ ϭ ͵ ␨ ͑␨͒ ͩ ͪ ͑ ͒ ˜ R p d G exp i ប⍀ p . 20 zation M/M 0 in the dHvA effect. Nonetheless, the first ob- Ϫϱ servation in 1930 of the SdH oscillations by Shubnikov and de Haas preceded the discovery of the dHvA effect.24 In the case of a spin ͑ ͒ Recently, Eq. 18 was proved experimentally in a 2D 1 organic superconductor of the ET family.7 The shape of the G͑␨͒ϭ ͓␦͑␨Ϫ␮ H͒ϩ␦͑␨ϩ␮ H͔͒, ͑21͒ 2 e e SdH oscillations observed was nonsinusoidal and very simi- ͑ ͒ ␮ ϭប lar to that given by Eq. 15 and shown in Fig. 1. What is where e e/2mec is the Bohr magneton. The correspond- strange in this connection that the SdH signal was measured s ing spin factor R p reads in a geometry in which both the magnetic field and current m were directed perpendicular to the layers. The temperature s ϭ ͩ ␲ ͪ ͑ ͒ R p cos p . 22 was nonzero and varied within the range 0.4–1.27 K. On the me other hand, it is easy to see that, strictly speaking, Eq. ͑18͒ This is a standard spin factor, known in magnetoos- does not hold under those conditions. First of all this is be- cillations.10 Another factor appears if impurities split off cause of the anisotropy of the conductivity in the direction bound states from the Landau levels.15 This magnetoimpurity perpendicular to the layers. Secondly, the layer factor I and p effect is stronger in 2D systems than in the 3D case. If we the factor RT ͑as well as some other factors which we discuss p denote this splitting by ⌬ (H), then the appropriate DOS is below͒ are functions of the inverse magnetic field 1/H, and i again given by the ␦ function .Hץ/ ˜Mץ this must be taken into account in the derivative i ␨͒ϭ␦ ␨Ϯ⌬ ͒ ͑ ͒ The additional degrees of freedom related to the spin of G ͑ ͑ i . 23 electron and bound states due to impurities ͑the so-called magnetoimpurity levels15͒ can be easily taken into consider- The corresponding damping factor due to the magnetoimpu- ation. In these cases the energy of an electron gains a shift rity bound states is ␧→␧ϩ␨ ␨ , where is an energy variable related to the addi- 2␲⌬ ␨ ␨ i ϭ ͩ i ͪ ͑ ͒ tional degree of freedom. If the DOS for is given by G( ), R p cos ប⍀ p . 24 then Ϯ ͑ ͒ → ͑ ͒ The signs in Eq. 23 correspond to attractive and repul- I p I pR p , 19 ⌬ Ӷប⍀ sive impurities. In the case i the splitting value does where not depend on p.16 Putting together all the above discussed Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Gvozdikov 707

damping factors and taking the factor I p in the form given by Eqs. ͑7͒–͑9͒, we can write the oscillating part of the in-plane conductivity in the form ␴ϭ␴ ϩ␴ ͑ ͒ ˜ ˜ 1 ˜ 2 25 with 2␲p ␴ ϭϪ ␴ ͑ Ϫ ͒ ͩ Ϫ ͪ ˜ 1 2 0 1 c ͚ exp ␶ ⍀ pϭ1 0 4␲t 2␲␮ ϫ T i s ͩ ͪ ͩ ͪ ͑ ͒ R pR pR pJ0 ប⍀ p cos ប⍀ p , 26

2␲p ␴ ϭϪ ␴ ͩ Ϫ ͪ ˜ 2 2 0c ͚ exp ␶ ⍀ pϭ1 1 4␲t 2␲␮ ϫ T i s ͩ ͪ ͩ ͪ ͑ ͒ R pR pR p cos ប⍀ p cos ប⍀ p . 27 ␯ ϭ ␲ ␶ ⍀ Here 1 2 / 1 , where ប ␶ ϭ ͑ ͒ 1 28 ϩͱ 2Ϫ 2 a t t0 is the renormalized ͑on account of the stacking faults͒ effec- tive electron lifetime. At zero temperature we have ␴ 0c ˜␴ ϭϪ 2 8 FIG. 2. Field dependence of the oscillating part of the conductivity caused sinh ␯ ϫ 1 Ϫ by the stacking faults and given by Eq. ͑29͒͑in conventional units B ͚ ␲ 1 . ϭប⍀͒ ␮ϭ ϭ ϭ ⌬ ϭ ϭ ␯ ϭ ␣,␤,␥ϭϮ ͭ m ͮ for 1000, c 0.1, t 0.1, i 0.5, m/me 1.5, 1 0.1, 0.2, 0.3 ␯ Ϫ ͫ͑␮ϩ␣⌬ ϩ␤ ͒ϩ␥ ͬ ␯ ϭ ␲ប ␶ cosh 1 cos i t ( 1 2 / 1). me ͑29͒ ␴ Despite the fact that ˜ 2 is proportional to the stacking fault ␯ Ӷ concentration c, it may be important at small 1 1, since in In this paper we have calculated the SdH oscillations ␴ ␦ this case ˜ 2 has a sharp -like peaks, as is seen in Fig. 2. within an approach which generalizes the Green function method developed by Maniv and Vagner23 for the electron scattering problem in a quasi-2D electron gas. The calculated CONCLUSIONS AND DISCUSSION oscillating part of the diagonal component of the in-plane The two major experimental tools of the fermiology, the conductivity tensor ˜␴, Eq. ͑13͒, and the electron lifetime ␶, SdH and the dHvA effects, have proved to be very useful not Eq. ͑11͒, turned out to be dependent on the very same layer ͑ ͒ only for studies of conventional 3D metals but also for recent factor I p , Eq. 1 , which modulates the dHvA oscillations of 12 analysis of the new layered organic conductors and super- the layered conductors and depends on the layer stacking. conductors, including the high-Tc cuprates. Though both ef- At zero temperature and in the 2D limit a simple relation was fects are based on the Landau quantization, the SdH effect, established between the relative in-plane SdH oscillations, 10,17,25 ␴ ␴ as is well known from the textbooks on fermiology, is ˜ / 0 , and the corresponding relative dHvA oscillations of ˜ ͑ ͒ much more difficult to study theoretically, since it requires the magnetization M/M 0 . This relation is given by Eq. 18 , calculations for the transition probabilities caused by elec- which holds only for the 2D electron gas. The anisotropy of tron scattering on impurities in an external magnetic field. the 2D Fermi surface and also nonzero temperature and elec- This scattering in turn strongly depends on the dimensional- tron hopping across the layers bring additional terms to Eq. ity of the system in question and manifests itself in different ͑18͒. Nonetheless, this equation has been experimentally kinetic characteristics of organic conductors, such as the verified by simultaneous dHvA and SdH measurements on 26 ␤ ͑ ͒ high-frequency impedance and the SdH in quasi-2D con- the organic conductor Љ– BEDT–TTF 2SF5CH2SO3 within ductors, which is the subject of the present publication. Con- the temperature range 0.4–1.27 K for the interlayer compo- trary to the wide-spread opinion, even among the experts, it nent of the conductivity tensor.7 The shape of these strongly is the SdH oscillations, rather than the dHvA effect, which nonsinusoidal oscillations is the same as given by Eq. ͑15͒ were first observed in 1930 in Leiden by Shubnikov and and shown in Fig. 1. The enhancement of the impurity scat- de Haas and thereby laid the foundation of modern fermiol- tering potential makes these oscillations lower in amplitude ogy. This point, as well as the dramatic history of the discov- but more sinelike in form. The spin splitting and magnetoim- s i ery of the SdH effect, can be found in the interesting histori- purity bound states yield additional factors, R p and R p , re- cal essay of Ref. 24. spectively, which modulate the SdH oscillations ͓see Eqs. 708 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Gvozdikov

The author is grateful to Tsofar Maniv and Israel Vagner for numerous stimulating discussions on the problem of elec- tron scattering in a quasi-2D gas in magnetic field and to Peter Wyder for hospitality during his stay at the Grenoble High Magnetic Field Lab.

*E-mail: [email protected]

1 F. A. Meyer, E. Steep, W. Biberacher, P. Crist, A. Lerf, A. G. M. Jansen, W. Joss, and P. Wyder, Europhys. Lett. 32, 681 ͑1995͒. 2 N. Harrison, J. Caulfeld, J. Singleton, P. H. P. Reinders, F. Herlach, W. Hayes, M. Kurmoo, and P. Day, J. Phys.: Condens. Matter 8, 5415 ͑1996͒. 3 E. Steep, L. H. Nguen, W. Biberacher, H. Muller, A. G. M. Jansen, and P. Wyder, Physica B 259–261, 1079 ͑1999͒. 4 M. M. Honold, N. Harrison, M. S. Nam, J. Singleton, C. H. Mielke, M. Kurmoo, and P. Day, Phys. Rev. B 58, 7560 ͑1998͒. 5 M. Shiller, W. Schmidt, E. Balthes, D. Schwitzer, H. J. Koo, M. H. Wan- hbo, I. Heinen, T. Kalusa, P. Kircher, and W. Strunz, Europhys. Lett. 51, 82 ͑2000͒. 6 M. V. Kartsovnik, G. Yu. Logvenov, T. Ishiguro, W. Biberacher, H. Anzai, and N. D. Kushch, Phys. Rev. Lett. 77, 2530 ͑1996͒. 7 J. Wosnitza, S. Wanka, J. Hagel, H. v. Lohneysen, J. S. Qualls, J. S. Brooks, E. Balthes, J. A. Schlueter, U. Geiser, J. Mohtasham, R. W. Win- ter, and G. R. Gard, Phys. Rev. Lett. 86, 508 ͑2001͒. 8 I. M. Lifshitz and A. M. Kosevich, Sov. Phys. JETP 2,639͑1953͒. 9 I. D. Vagner, T. Maniv, and E. Ehrenfreund, Phys. Rev. B 51,1700͑1983͒; K. Jauregui, V. I. Marchenko, and I. D. Vagner, Phys. Rev. B 41, 12922 ͑1990͒; P. D. Grigor’ev and I. D. Vagner, JETP Lett. 69,156͑1999͒. 10 D. Shoenberg, Magnetic Oscillations in Metals, Cambridge University Press, Cambridge ͑1984͒. 11 T. Champel and V. P. Mineev, Philos. Mag. B 81,55͑2001͒. 12 ͑ ͒ ͑ ͓͒ FIG. 3. Field dependence of the oscillating part of the conductivity given by V. M. Gvozdikov, Fiz. Tverd. Tela Leningrad 26, 2574 1984 Sov. ͑ ͔͒ Eq. ͑25͒–͑28͒͑in conventional units Bϭប⍀͒ for ␮ϭ1000, cϭ0.2, ⌬ Phys. Solid State 26, 1560 1984 . i 13 ͑ ͒ ͑ ͓͒ ϭ0.5, m/m ϭ1.5, Tϭ0.001, ␯ϭ0.1, ␯ ϭ0.2, tϭ0.1, 0.3, 0.5 ͑␯ V. M. Gvozdikov, Fiz. Tverd. Tela Leningrad 28, 320 1984 Sov. Phys. e 1 ͑ ͔͒ ϭ2␲ប/␶ , ␯ ϭ2␲ប/␶ ͒. Solid State 28, 179 1986 . 0 1 1 14 V. M. Gvozdikov, Fiz. Nizk. Temp. 12, 705 ͑1986͓͒Sov. J. Low Temp. Phys. 12, 399 ͑1986͔͒. 15 ͑ ͒ ͑ ͒ ͑ ͔͒ A. M. Ermolaev and M. I. Kaganov, JETP Lett. 6, 171 1967 . 25 – 27 and result in a large variety of oscillation pat- 16 V. M. Gvozdikov, A. M. Ermolaev, and Yu. A. Sharapov, Vestnik Khark- terns, as displayed in Figs. 2 and 3. ovskogo Natsional’nogo Universiteta, 476, Ser. ‘‘Fizika,’’ N 4, 18 ͑2000͒. The stacking faults modify the Dingle-like exponent in 17 A. A. Abrikosov, Fundamentals of Theory of Metals ͑in Russian͒ Nauka, ͑ ͒ Eq. ͑27͒ through the effective lifetime ␶ ͓Eq. ͑28͔͒. The Moskow 1987 . 1 18 T. Ando, A. B. Fowler, and F. Stern, Rev. Mod. Phys. 54,437͑1982͒. ␴ 19 appropriate correction to the oscillation ˜ 2 arises due to the N. Harrison, R. Bogaerts, P. H. P. Reinders, J. Singleton, S. J. Blundell, stacking faults and at zero temperature has periodic ␦-like and F. Herlach, Phys. Rev. B 54, 9977 ͑1996͒. ␯ Ӷ 20 A. E. Datars and J. E. Sipe, Phys. Rev. B 51, 4312 ͑1995͒. peaks when 1 1. It is substantial even for low concentra- 21 R. H. Mc Kenzie and P. Moses, Phys. Rev. Lett. 81, 44922 ͑1998͒. tions of stacking faults, as one can see in Fig. 2. 22 T. Ando and Y. Uemura, J. Phys. Soc. Jpn. 36, 959 ͑1974͒. The oscillation patterns displayed in Figs. 1–3 are very 23 T. Maniv and I. D. Vagner, Phys. Rev. B 38, 6301 ͑1988͒. diverse. The peaks in Figs. 2 and 3 are split because of the 24 B. I. Verkin, S. A. Gredeskul, L. I. Pastur, and Yu. A. Freyman, Fiz. Nizk. spin and magnetoimpurity factors and are sensitive to varia- Temp. 16, 1203 ͑1990͓͒Sov. J. Low Temp. Phys. 16, 696 ͑1990͔͒. 25 I. M. Lifshits, M. Ya. Azbel’, and M. I. Kaganov, The Electron Theory of tions of the other parameters of the model. Some additional Metals, Nauka, Moskow ͑1971͒. analysis of the SdH oscillations within this model will be 26 V. G. Peschansky, I. V. Kozlov, and K. Jiasemides, Fiz. Nizk. Temp. 26, published elsewhere. The results and methods of this paper- 225 ͑2000͓͒Low Temp. Phys. 26, 169 ͑2000͔͒. will be also applied to the anisotropic 2D Fermi-surface This article was published in English in the original Russian journal. Repro- structures typical for real organic layered ͑super͒ conductors. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Giant magnetooscillations of the Josephson current L. P. Gor’kov

National High Magnetic Field Laboratory, Florida State University, Tallahassee, FL 32306, USA; L. D. Landau Institute for Theoretical Physics, Russia V. Z. Kresin*

Lawrence Berkeley Laboratory, University of California, Berkeley, CA 94720 ͑Submitted June 19, 2001͒ Fiz. Nizk. Temp. 27, 963–966 ͑September–October 2001͒ Josephson current flowing through an antiferromagnetic metallic barrier is studied. We consider the most interesting case of the spin-valve structure when the barrier is formed by ferromagnetic layers ordered antiferromagnetically in the direction perpendicular to the current. The most remarkable feature is the onset of giant magnetooscillations of the current amplitude. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1401179͔

INTRODUCTION The low-temperature phase diagram has been analyzed in our previous papers.2,3 The parent material, LaMnO ,isan The study of various resonance phenomena is an impor- 3 insulator. The region 0.17ϽxϽ0.5 corresponds to a metallic tant area of condensed matter physics related to unique in- ͑ ͒ state. The undoped manganite e.g., LaMnO3 crystal has the formation about the spectra and properties of various mate- ϩ following cubic structure. The Mn3 ions are located at the rials. Measurements of such phenomena as Shubnikov-de corners, and the La ion is at the center of the unit cell. In Haas oscillations, de Haas-van Alphen effects, cyclotron ϩ Ϫ addition, the Mn3 is caged by an O2 octahedron; locally resonance, etc., are well known and are described in many this forms an MnO complex with the Mn ion in the central textbooks and monographs. 6 position surrounded by light O ions. In this article we will describe the giant magnetooscilla- The d shell of the Mn ion in the cubic environment is tion effect. The proposed method, which combines Joseph- split into a doublet and triplet. It is important that the three- son tunneling spectroscopy and novel magnetic systems, is fold manifold (t ) is occupied by three d electrons, whereas very sensitive to the magnetic structure of compounds. 2g the upper doubly degenerate term e is occupied by one We focus on the special case of the S–N–S junctions 2g electron only. when the normal ͑N͒ metallic barrier is a magnetic com- The key ingredients ͑interactions͒ are the following: 1͒ pound. It is well known that the ferromagnetic material ͑F͒ the strong Hund’s rule coupling ͑this is the largest energy used as a barrier frustrates the Josephson current because of scale͒ aligns all spins in the same direction; 2͒ hopping, and the pair-breaking effect of the exchange field. Below we con- 3͒ cooperative Jahn–Teller ͑JT͒ effect. As a result, the total sider an antiferromagnetic barrier and, more specifically, the ͑ Hamiltonian is a sum: most interesting case of the so-called A-structure S1–A–S2 junction͒. This is the situation when the barrier is formed by ϭ ϩ ϩ ͑ ͒ a multilayer system. It consists of ferromagnetic layers per- H HH Ht HJT . 1 pendicular to the S electrodes, with an antiferromagnetic or- dering in the direction perpendicular to the current. In other Here words, neighboring layers have opposite magnetization. ϭ ␴ ͑ ͒ At first, we describe the compounds and their A struc- HH JH͚ •Si , 2 ture. Afterwards, we introduce the giant magnetooscillation i phenomenon for the S1–A–S2 junction. ϭ ͑ ͒ Ht ͚ ti,iϩ␦ , 3 i A-STRUCTURE. SPIN-VALVE EFFECT ϭ ͑␶ ͒ϩ 2 ͑ ͒ There are two interesting systems which are character- HJT ͚ g i Qi ͚ JelQi , 4 i • i ized by A-structure. One of them is a mixed-valence manga- nite with some characteristic level of doping, and the second where Q is a normal coordinate, ␴ is the electronic spin, S is ␶ one is an artificial structure, a so-called giant magnetoresis- the ionic spin, Jel is the elastic parameter, and is the pseu- tance ͑GMR͒ multilayer. dospin. As is known, the most remarkable property of mangan- The Hund’s coupling between the local spin S (S ͑ 1͒ ϭ ites see review is the colossal magnetoresistance. These 3/2) formed by the t2g electrons and the e2g electron is ϭ ͑ ͒ ␴ ͑ ͒ materials with the composition A1ϪxSrxMnO3 (A La, Ca) described by the term 2 ;. Pauli matrices correspond to ϳ have a rather complicated phase diagram. the spin of the e2g electron; JH 1 eV. Note that the unit cell

1063-777X/2001/27(9–10)/4/$20.00709 © 2001 American Institute of Physics 710 Low Temp. Phys. 27 (99–10), September–October 2001 L. P. Gor’kov and V. Z. Kresin contains one e2g electron, and its motion through the lattice magnetic and nonmagnetic layers, with an antiferromagnetic is described ͑in the tight-binding picture͒ by the term ͑3͒. ordering in the direction perpendicular to the layers. The third term in Eq. ͑1͒ describes another important Let us evaluate the energy spectrum for such a metallic ingredient which also affects the behavior of the system, A-system. This is an important step for the calculation of the namely the Jahn–Teller instability. Indeed, the e2g electron is Josephson current through such a barrier. For concreteness, in the doubly degenerate state, and it follows from the JT we consider a metallic manganite in the A-phase. theorem that the electron–lattice coupling will lead to a static The Hamiltonian for the barrier’s states has the form: distortion and consequently, to a change in the crystal sym- metry. In Eq. ͑4͒ g is the electron–phonon coupling constant, ϭ ͑ ͒ ϩ ϩ ϩ ͑ ͒ ϩ ͑␴ ͒ H ͚ t p ap␴ap␴Ј ͚ JHS Q ap␴Ј ˆ z ␴Ј␴ЉapϪQ;␴Љ and Qi are the local active JT modes. p p,Q In a simple band picture with one electron per unit cell, the system should be metallic. This is contrary to the experi- ϩ ͑Ϫ ͒ ϩ ͑␴ ͒ ͚ JHS Q ap␴Ј ˆ z ␴Ј␴ЉapϪQ;␴Љ mental fact that the parent compound is an insulator. Never- p,Q theless, a more careful analysis which includes not only the ϩ ϩ ͑␴ ͒ ͑ ͒ hopping term, but also a strong Hund’s interaction along ͚ JHMap␴Ј ˆ x ␴Ј␴ЉapϪQ;␴Љ . 6 with the cooperative Jahn–Teller effect, leads one to the pic- p 2,3 ture of a peculiar band insulator. Here tϭtЌϩtʈ , tЌ and tʈ are electron hopping param- As noted above, Hund’s coupling corresponds to the eters for the in-plane and out-of-plane motion, J is the ϳ ӷ H largest energy scale (JH 1 eV), so that JH t, gQ0 ; t Hund’s coupling, and S(Q) is the Fourier component of the ϳgQ ϳ0.1 eV. ϭ Ϫ n 0 AFM ordering along the c directions; ͗SZ͘ S( 1) . The It is interesting that the magnetic structure of the un- structural vector Qϭ(0,0,␲/a) reduces the Brillouin zone ͑a doped crystal belong to the A-type: antiferromagnetic ͑AF͒ is the lattice constant͒. We consider the more general case of ordering along the Z axis along with the ferromagnetic or- a canted structure; M is the canted magnetic moment: Si dering in the XY plane. The appearance of such a structure is ϭ Ϯ͗ ͘ 2ϩ 2ϭ ϩ Х 2 ( Sz ,M x), Sz M x S(S 1) S . Note that the caused by an interplay between the Hund’s interaction and Jahn–Teller effect is not essential for the metallic state. the hopping energy. In other words, the A-structure is stabi- With the use of the equation of motion lized by the gain in the kinetic energy of the band electron. One can explain the underlying physics by considering the ϩ ϩ ͑ 3 Ϫ 4 ͒ ͑␧Ϫ ʈ͒ ϭ ͑Ϯ ͒␴ ϩ ͑␴ ͒ two-center problem Mn Mn with spins S1 and S2 . ˜ t ak␴ JHS Q ˆ zakϪQ;␴ JHM ˆ x ␴␴Јak␴Ј The low-energy electronic term has the form: ͑7͒

and a similar equation for k→kϩQ (˜␧ϭ␧ϪtЌ), one can ϭϪ ϩ Ϫ 2 ͒ ͑ ͒ E JHS t cos q/2 O͑t /JH . 5 determine the following energy spectrum:

Therefore, at J ӷt, the main gain in the energy is due to the 2 2 2 1/2 H ˜␧ϭϮ͓J S Ϯ2J Mtʈϩtʈ ͔ ͑8͒ Ϫ H H JHS term. Although, because of the averaging, there is no ӷ linear contribution to the total energy, a small gain, of the which consists of four branches. For the manganites JH t 2 2 Ӷ ͑ ͒ order of t /JH , is achieved. The relation t /JH t sets a see above . As a result, for these systems, only two branches lower energy scale for the Ne´el temperature TN , which is are filled: indeed small ͑ϳ140 K͒ relative to the structural transition temperature ͑ϳ0.1 eV͒. Within a single layer the electrons ˜␧ ХϪJ SϮ͑M/S͒tʈ . ͑9͒ may be treated as ferromagnetically polarized. This ferro- 1,2 H magnetism is provided by the double-exchange mechanism introduced by Zener4 and developed in Refs. 5 and 6. JOSEPHSON CURRENT THROUGH AN ANTI- FERROMAGNETIC BARRIER. MAGNETOOSCILLATIONS As was noted above, the pure LaMnO3 compound can be treated as a band insulator. This means that a change in the Let us evaluate the Josephson current through the carrier concentration might lead to metallic behavior. A-barrier. We consider the case of a singlet pairing ͑s or d͒, Consider now a doped manganite, e.g., La1ϪxCaxMnO3. i.e., the Cooper pair consists of two carriers with opposite An increase in x will lead initially to the formation of finite spins and momenta. The A-barrier contains states with oppo- clusters; each of them contains itinerant polarized electrons. site spins; they belong to neighboring layers and, contrary to Eventually one can observe the formation of an infinite clus- the F barrier, one should not expect any frustration for the ter and the transition to the metallic ferromagnetic state. Josephson current. Nevertheless, the calculation of the cur- Such transition can be described by percolation theory. rent should be carried out with considerable care. Indeed, the A very interesting observation was reported in Refs. Josephson current is a transfer of correlated electrons, and 7–10. The presence of a metallic A-phase was observed. The this implies that the layers are not totally independent. The ϭ examples are the compounds A1ϪxSrxMnO3 (A La, Nd) barrier should be treated as an anisotropic metal with the ϭ with x 0.55, or Pr0.5Sr0.5MnO3. This is a natural spin-valve spectrum ͑8͒, ͑9͒. The Josephson current can be evaluated structure. with use of the spectrum ͑8͒, ͑9͒ and the interface Hamil- Another case is an artificial A-structure, which is basic to tonian: the GMR phenomenon ͑see reviews11,12͒. Such a multilayer ͑ ͒ ϭ ⌬ ⌿ ͒⌿ϩ ͒ ͑ ͒ system e.g., Co–Al2O3–Co consists of alternating ferro- Hpair V i ͑i ͑i 10 Low Temp. Phys. 27 (99–10), September–October 2001 L. P. Gor’kov and V. Z. Kresin 711 which describes the transition of a pair in the ith supercon- canting is small, and the dependence of the amplitude is ϭ ͑⌬ ⌬ ␤ ͑ ͒ ␤ ductor (i 1,2) into the barrier 1 , 2 are the order param- determined by the factor J0( M/S) in Eq. 14 . Since eters, and ⌿(i) and ⌿ϩ(i) are the field operators for elec- ӷ1, one can use an asymptotic expression for the Bessel trons of the barrier͒. With the use of Eq. ͑10͒ we find a function, and we obtain: correction to the thermodynamic potential, ␦⍀, caused by the L barrier; the current is obtained as ␦⍀/␦␸, with ␸ being a ϭ ͓ Ϫ͑ ͒2͔ ͩ Ϫ ͪ Jm r 1 M/S exp ␰ ͑ ͒ phase difference between two superconductors’ order param- N T eters. The Josephson current is determined by the Cooper ␲␤M Ϫ1/2 ␤M ␲ diagram: ϫͩ ͪ cosͩ Ϫ ͪ . ͑15͒ 2S S 4 ϭ␲ ͉ 2͉ ⌬ ⌬ ͵ ͑ ͒⌸͑ ␻ ͒ The expression is valid if Lӷ␨ . K T V ͚ 1 2 dpdq exp iq L i nЈq , N ␻ • n Equations ͑14͒ and ͑15͒ display a remarkable property of ͑ ͒ 11 a junction with an A-barrier, namely, the phenomenon of where giant magnetooscillations ͑GMO͒. Indeed, one can see di- rectly from Eq. ͑15͒ that the amplitude oscillates as a func- ⌸͑ ␻ ͒ϭ ͑␴ ͒ ͑ ␻ ͒͑␴Ј͒ tion of M, and, therefore, as a function of the external field i n ,q ͚ ˆ y ␴,␴ЈG␴,␴Ј i n ,p y ␴Љ,␴ٞ .␴,␴Ј,␴Љ,␴ٞ H The GMO effect is caused by magnetic structure ͑spin- ϫG␴ ␴ ͑Ϫi␻ ,qϪp͒ -Ј, ٞ n valve effect͒ and is entirely different from the usual Fraun ϭ ␻ ͒ Ϫ ␻ Ϫ ͒ ␤ӷ 2͓G↓↑͑i n ,p G↑↓͑ i n ,q p hofer oscillations. Note also that, because 1, the period of the GMO is much smaller. Moreover, one can choose a ge- Ϫ ͑ ␻ ͒ ͑Ϫ ␻ Ϫ ͔͒ ͑ ͒ G↓↓ i n ,p G↑↑ i n ,q p 12 ometry of the Josephson contact such that the only the mag- ͑we use the method of thermodynamic Green’s functions; netic structure is affected by the magnetic field. An increase see, e.g., Ref. 13͒, and L is the distance between the super- in magnetic field eventually leads to the situation that M ϭ ϭ ͑ ͑ ͒ ͑ ͒͒ conductors. S; then Jm 0 see Eqs. 14 and 15 . This occurs be- The Green’s function appearing in Eq. ͑12͒ can be ex- cause the external field affects the magnetic ordering in the pressed in terms of new Fermi amplitudes which correspond barrier. If MϭS, then we are dealing with 3D ferromagnetic to the branches ͑8͒, ͑9͒. The transformation to the new am- ordering. As to the the S–F contact, it frustrates the plitudes can be described by a canonical transformation Josephson current; this is due to the pair-breaking effect of ϭ⌺ ␣ ␴ϵ ↑↓ ϭ ͑ ap␴ iK␴i ip ; one can show that ( ), K␴1 0.5͓1 the exchange field we are not considering some modifica- Ϫ 1/2 ϭϪ ϭ ϩ 1/2 ͑ ͒ (M/S)͔ ; K↑2 K↓2 0.5͓1 (M/S)͔ , where the tions caused by the spin–orbital coupling; see Ref. 14 . ␣ ͑ ͒ ͑ ͒ operators ip correspond to the branches 8 , 9 . Then we Therefore, the amplitude of the current can be controlled by obtain the following expression for the Cooper term: an external magnetic field and even switched off at larger H ͑order of 1 T͒. Note that for the GMR structure the switching M ϭͩ Ϫ ͪ ͉ 2͉͉⌬ ͉2͉⌬ ͉2␲ can be obtained with the use of a weaker magnetic field ͑of jm 1 V 1 2 T S the order of 3–4 Oe͒. In summary, we have shown that Josephson current can ␻ L LtʈM ϫ ͵ ͵ ͩ Ϫ n ͪ ͩ ͪ ͚ dlvЌ dpz exp cos . flow through an antiferromagnetic barrier. The most interest- ␻ Ͼ vЌ SvЌ n 0 ing case corresponds to the case of A-ordering in the direc- ͑13͒ tion perpendicular to the current. One can show that the am- plitude of the current oscillates as a function of applied In the tight-binding approximation tʈϭt cos(p d ). Equation 0 z c magnetic field ͑GMO͒. ͑13͒ contains the integration over the cross section of the Lev Vasilievich Shubnikov was a remarkable scientist Fermi surface ͑dl͒. who made key contributions to various areas of physics. We As a result, we arrive at the following expression for the are pleased that this paper is published in the Special Issue amplitude of the Josephson current: devoted his memory. ϭ Ϫ ͒2 ␨ ͒ ␤ ͒ ͑ ͒ Jm r͓1 ͑M/S ͔ f ͑L/ N J0͑ M/S . 14 ␨ ϭ ␲ Here N hvF/2 T is the coherence length inside the bar- ␤ϭ ␰ ␰ ϭ ␲ ϭ 0 *E-mail: [email protected] rier, (t0 /Tc)(L/ 0); 0 hvF/2 Tc , vF vF is the mini- mum value of the component of the Fermi velocity along L, ϰ ⌬ ⌬ 2 ϭ Ϫ ӷ ϭ Ϫ1 r p 1 2V , and f (x) exp( x) for x 1 and f (x) x Ӷ ϭ ϭ␨ for x 1. Therefore, near T 0 K, the function f N /L. 1 ϰ J. Coey, M. Viret, and S. von Molnar, Adv. Phys. 48, 167 ͑1999͒. Note that, since r T, the amplitude Jm in this case does not 2 L. Gor’kov and V. Kresin, JETP Lett. 67,985͑1998͒. 3 depend on T. However, near Tc and in the intermediate tem- M. Dzero, L. Gor’kov, and V. Kresin, Eur. Phys. J. B 14,459͑2000͒. perature region, the exponential dependence f ϭexp(ϪL/␨ ) 4 G. Zener, Phys. Rev. 82, 403 ͑1951͒. N 5 is perfectly realistic. P. Anderson and H. Hasegawa, Phys. Rev. 100,675͑1955͒. 6 P. de Gennes, Phys. Rev. 118,141͑1960͒. 7 H. Kawano, R. Kajimoro, H. Yoshizawa, Y. Tomioka, H. Kuwahara, and DISCUSSION Y. Tokura, Phys. Rev. Lett. 78, 44253 ͑1997͒. 8 H. Fujisturo, M. Ikebe, and Y. Konno, J. Phys. Soc. Jpn. 67, 1799 ͑1998͒. Consider the junction placed in an external magnetic 9 M. Isumi, T. Manako, Y. Konishi, M. Kawasaki, and Y. Tokura, Phys. Rev. field, which creates the canting. For weak magnetic field the B 61, 12187 ͑2000͒. 712 Low Temp. Phys. 27 (99–10), September–October 2001 L. P. Gor’kov and V. Z. Kresin

10 Y. Morimoto, T. Akimoto, A. Nakamura, K. Ohoyama, and M. Ohashi, Field Theory in Statistical Physics, Dover, New York ͑1975͒. Phys. Rev. B 58, 5544 ͑1998͒. 14 E. Demler, G. Arnold, and M. Beasley, Phys. Rev. B 55, 15174 ͑1997͒. 11 S. S. P. Parkin, Annu. Rev. Mater. Sci. 25, 358 ͑1995͒. 12 Ultrathin Magnetic Structures II, B. Heinrich and J. A. C. Bland ͑Eds.͒, Springer-Verlag, Berlin ͑1994͒. This article was published in English in the original Russian journal. Repro- 13 A. Abrikosov, L. Gor’kov, and I. Dzyaloshinskii, Methods of Quantum duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

The critical current of superconductors: an historical review D. Dew-Hughes*

Oxford University, Department of Engineering Science, Parks Road, Oxford OX1 3PJ, UK ͑Submitted May 4, 2001͒ Fiz. Nizk. Temp. 27, 967–979 ͑September–October 2001͒ The most important practical characteristic of a superconductor is its critical current density. This article traces the history of the experimental discoveries and of the development of the theoretical ideas that have lead to the understanding of those factors that control critical current densities. These include Silsbee’s hypothesis, the Meissner effect, the London, Ginsburg–Landau, and Abrikosov theories, flux pinning and the critical state, and the control of texture in high-temperature superconductors. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1401180͔

INTRODUCTION THE EARLY YEARS 1911–1936

The most important characteristic of any superconductor, Within two years of his discovery of superconductivity from the viewpoint of practical applications, is the maximum in mercury, Onnes recorded that there was a ‘‘threshold electrical transport current density that the superconductor is value’’ of the current density in mercury, above which the able to maintain without resistance. This statement is equally resistanceless state disappeared.3 This critical value was tem- true for large-scale applications, such as power transmission perature dependent, increasing as the temperature was re- lines, electromagnets, transformers, fault-current limiters and duced below the critical temperature, according to the rotating machines, as well as for small-scale electronic ap- expression4 plications such as passive microwave devices and devices ͒ϭ ͒ Ϫ ͒ ͑ ͒ based on the Josephson effect. High lossless current densities Jc͑T Jc͑0 ͑Tc T /Tc . 1 mean that machines and devices can be made much smaller A similar behavior was observed in small coils fabricated and more efficient than if made with a conventional resistive 5 conductor. This was realized immediately upon the discovery from wires of tin and lead. These represent the first-ever of superconductivity; Onnes himself speculating on the pos- superconducting solenoids. Also noticed was the fact that the sibility of magnet coils capable of generating fields of critical current density in the coils was less than that ob- 105 G.1 These early hopes were dashed by the inability of the served in short, straight samples of wire. This is the first then-known superconductors to sustain substantial currents, instance of the phenomenon that was to plague the designers and applied superconductivity did not become a commercial of superconducting magnets. reality until alloy and compound superconductors based on The following year Onnes reported on the influence of a the element niobium were developed around 1960.2 In the magnetic field on the superconducting transition in lead: ‘‘The introduction of the magnetic field has the same effect two following decades, intensive effort, primarily by metal- 6 lurgists, led to the understanding of the factors that control as heating the conductor.’’ The existence of a critical mag- critical currents and to the development of techniques for the netic field, above which superconductivity ceased to exist, fabrication of complex multifilamentary flexible conductors was demonstrated. Surprisingly, perhaps because of the in- at economic prices. The discovery of the mixed copper oxide tervention of the First World War, Onnes failed entirely to high-temperature superconductors initially produced a disap- make the connection between the critical current and the pointment similar to that experienced by the pioneers of su- critical magnetic field. This connection was left to be made perconductivity. The superconducting characteristics of these by Silsbee, as a consequence of his examining all of Onnes’ materials introduced a new set of obstacles to achieving cur- published reports in great detail. Silsbee’s hypothesis states, rent densities of magnitudes sufficient for practical device ‘‘The threshold value of the current is that at which the mag- netic field due to the current itself is equal to the critical applications. The difficulties involved in producing long 7 lengths of high-current conductor from these materials are magnetic field.’’ From outside a conductor of circular cross- only just being overcome. section, carrying a current I, the current appears to flow in a This article is not intended to be a review of everything dimensionless line down the middle of the conductor. At a that is known about critical currents in superconductors. Its distance r away from a line current, there is a tangential aim is to trace the historical development of the understand- magnetic field of strength ing of the factors that control critical current density in su- H͑r͒ϭI/2␲r. ͑2͒ perconductors. The significant experimental facts and theo- retical ideas that have contributed to the present level of If the radius of the conductor is a, then the field at the surface knowledge will be outlined, and the crucial contribution to of the conductor will be the topic made by Lev Vasilievich Shubnikov will be high- lighted. H͑a͒ϭI/2␲a ͑3͒

1063-777X/2001/27(9–10)/10/$20.00713 © 2001 American Institute of Physics 714 Low Temp. Phys. 27 (99–10), September–October 2001 D. Dew-Hughes and the critical current, according to Silsbee’s hypothesis, transition temperature to well above that of the supercon- will be ducting element, but these alloys also exhibited very high ϭ ␲ ͑ ͒ critical fields. These investigations culminated in the discov- Ic 2 aHc . 4 ery that the Pb–Bi eutectic had a critical field of about 20 kG It should be noted that the critical current is thus not an at 4.2 K, and its use to generate high magnetic fields was 13 intrinsic property of a superconductor, but is dependent upon proposed. This was actually attempted at the Clarendon the size of the conductor, increasing as the diameter of the laboratory in Oxford, to which Lindemann had recruited Si- conductor is increased. Conversely, the critical current den- mon, Kurti, and Mendelssohn as refugees from Nazi Ger- sity, also size dependent, decreases as the diameter of the many. The attempt failed, as did a similar one by Keesom in conductor is increased: the Netherlands. Resistance was restored at levels of mag- netic field more appropriate to pure elemental superconduct- ϭ ͑ ͒ Jc 2Hc /a. 5 ors. The conclusion was that the Silsbee’s hypothesis was not 14 The experimental confirmation of Silsbee’s hypothesis valid for alloys. had to wait until after the end of the war. Both Silsbee8 and The studies on tin single crystals at Leiden had produced the Leiden laboratory9 carried out experiments on wires of the puzzling results that, in a transverse field, resistance was restored at a value of field one-half of the critical field when differing diameters that did indeed confirm the correctness of 15 the hypothesis. Tuyn and Onnes stated, ‘‘On the faith of the field was applied parallel to the axis of the crystal. Von these results obtained up till now we think we may accept the Laue, better known for his x-ray work, realized the signifi- hypothesis of Silsbee as being correct.’’ Silsbee’s summary cance of this result and suggested that it would be profitable to explore the distribution of magnetic field in the neighbor- was, ‘‘It may therefore be concluded that the results of these 16 experiments can be completely accounted for by the assump- hood of a superconductor. tion of a critical magnetic field, without making use of the Meissner had already interested himself in this problem; concept of critical currents.’’ he and others had considered the possibility of a supercurrent Equation ͑2͒ is valid whatever the actual distribution of being essentially a surface current. In 1933 Meissner and the current inside the conductor, and therefore Eq. ͑4͒ also Ochsenfeld published the results of their experiments in holds for a hollow conductor of the same external radius. An which they measured the magnetic field between two parallel ingenious extension of the Leiden experiments was to mea- superconducting cylinders. The enhancement of the field as sure the critical current of a hollow conductor in the form of the temperature was lowered below the critical temperature of the cylinders indicated that flux was being expelled from a film of tin deposited on a glass tube. An independent cur- 17 rent was passed along a metal wire threaded through the the body of the superconductors. tube. Depending on the direction of this current the critical Shubnikov had left Leiden in 1930 to take up a position current of the tin was either augmented or decreased, as the at the Ukrainian Physicotechnical Institute in Kharkov, field at its surface resulted from both currents in the tin film where he shortly became the scientific director of the newly and in the wire. This reinforced the validity of Silsbee’s hy- established cryogenic laboratory. Liquid helium became pothesis. available in the laboratory in 1933, and in the following year Rjabinin and Shubnikov gave confirmation of the Meissner At the same time, the Leiden laboratory was also making 18 a study of the temperature dependence of the critical field in effect in a rod of polycrystalline lead. tin, with the result:9,10 The importance of this discovery of the Meissner effect to the understanding of superconductivity cannot be overem- T 2 phasised. A perfect conductor will exclude flux if placed in ͑ ͒ϭ ͑ ͒ͫ Ϫͩ ͪ ͬ ͑ ͒ Hc T Hc 0 1 . 6 Tc an increasing magnetic field, but should retain flux if cooled to below its transition temperature in a magnetic field. The Also hysteresis in the superconducting transition was ob- Meissner effect is the expulsion of flux from the body of a 11 served for the first time. Hysteresis was subsequently ob- superconductor when in the superconducting state. The tran- served in indium, lead, and thallium, and it was suggested sition from the normal state to the superconducting state is that it might be an effect of purity, strain, or crystalline path independent, and the superconducting state is thermo- 12 inhomogeneity. It was decided that measurements on single dynamically stable. Armed with this knowledge it was pos- crystals would be desirable, and in 1926 Shubnikov, who at sible to develop phenomenological theories of superconduc- that time was an expert in the growth of single crystals, tivity. Being the more stable state below the transition joined the Leiden laboratory on a four-year secondment. temperature, the superconducting state has a lower energy Meanwhile, in 1925, a new liquid helium laboratory was than the normal state. It is possible to show, from simple established at the Physikalische Technische Reichsanstalt in thermodynamics, that the energy per unit volume of the su- Charlottenberg. Chosen as the head of this laboratory was a perconducting state relative to the normal state is former student of Planck, Walther Meissner. Meissner imme- diately instituted a program of work on superconductivity, 1 ⌬G ϭϪ ␮ H2. ͑7͒ but in order to avoid conflict with the Leiden group, this ns 2 0 c program concentrated on the superconducting transition metals, in particular tantalum and niobium. This is in fact just the energy required to exclude the mag- At Leiden attention had now turned to binary alloys, one netic field from the superconductor. constituent of which was a superconductor and the other a Two phenomenological theories followed almost imme- nonsuperconductor. Not only did alloying often raise the diately from the discovery of the Meissner effect. The ‘‘two- Low Temp. Phys. 27 (99–10), September–October 2001 D. Dew-Hughes 715

fluids’’ model of Gorter and Casimir19 was able to describe not specified, but it was assumed that the meshes were of a the influence of temperature on the properties of the super- dimension small compared to the penetration depth. Gorter25 conducting state, and it is similar to the theory for liquid produced an alternative proposal, that the alloy supercon- helium below its lambda point. In particular, the temperature ductors subdivided into extremely thin regions, rather like a dependence of the critical magnetic field, Eq. ͑6͒, can be stack of razor blades, parallel to the applied field. This sug- derived from the two-fluid model. The London theory deals gestion is remarkable in the light of Goodman’s lamellar with the effect of magnetic fields upon the superconducting theory for type-II superconductivity.26 However, even more properties, and describes the spatial distribution of fields and prescient was Gorter’s notion of a minimum size for the currents within a superconductor.20 The Londons showed superconducting regions, foretelling the later concept of the that flux was not totally excluded from the body of a super- coherence length. conductor, but that it penetrated exponentially, from the sur- Because in an ideal superconductor the flux expulsion is face, decaying over a characteristic length ␭, the penetration not complete, some surface penetration occurring, the energy depth required to expel the flux is less than that given by Eq. ͑7͒, and the actual critical field is slightly higher than that pre- ͑ ͒ϭ ͑ ͒ ͑Ϫ ␭͒ ͑ ͒ H r H 0 exp r/ . 8 dicted from complete expulsion. This effect is barely notice- Associated with the gradient in field is a current able in bulk superconductors, but can become appreciable when at least one dimension of the superconductor is com- .H H͑0͒ parable to, or smaller than, the penetration depth. Hץ J͑r͒ϭ ϭϪ exp͑Ϫr/␭͒. ͑9͒ 27 -r ␭ London showed that the critical field for a slab of superץ conductor, of thickness d, in an external field parallel to the Note that this current has a maximum value at the surface, faces of the slab is given by ϭ ␭ r 0, equal to Hc(0)/ . This is the maximum current den- Ϫ1/2 sity that a superconductor can tolerate, and for lead, for ex- ␭ d H ϭH ͭ 1Ϫ tanh ͮ . ͑10͒ ample, with a critical field at 4.2 K of ϳ4.2ϫ104 A/m and a f c d ␭ penetration depth of ϳ35 nm, this maximum current density ␭ is ϳ1.2ϫ1012 A/m2. Another important result of the London When d is small compared to , this reduces to theory was the conclusion that magnetic flux trapped by ␭ holes in a multiply connected superconductor, or within the H ϭ) H . ͑11͒ f d c body of the superconductor, must be quantized. The quantum ⌽ ϭ of magnetic flux was shown to be 0 h/q, where h is Thus thin films can remain superconducting to higher fields, Planck’s constant and q is the charge of the carrier associated and carry higher currents, than can bulk superconductors. with superconductivity. This suggestion was verified experimentally by Shalnikov in The groups at Oxford, Leiden, and Kharkov continued 1938.28 London suggested that, if the surface energy between their studies on alloys. The addition of 4% Bi to Pb was normal and superconducting regions was negative, the super- sufficient to completely trap magnetic flux when an external conductor would split into alternate lamellae of normal and field was reduced from above the critical field to zero.21 In superconducting regions, as suggested by Gorter. Fine fila- alloys of Pb–Ti and Bi–Ti, in increasing applied fields, flux ments, of diameter less than the coherence length, are ex- began to penetrate at fields well below those at which resis- pected, by similar arguments, to have a higher critical field tance was restored.22 Rjabinin and Shubnikov’s work on than that of the bulk. The Mendelssohn sponge could well be single crystals of PbTl2 clearly demonstrated the existence of a mesh of fine filaments, with superconducting properties ͑ two critical fields. Below the lower critical field, Hk1 in slightly better than those of the matrix. The filaments are their notation͒, the alloy behaved as a pure metal supercon- assumed to result from inhomogeneities in the two-phase ductor, with no flux penetration. Above Hk1 flux began to Pb–Bi alloys under investigation. penetrate; penetration was completed at the upper critical The picture emerging by mid-1935 was that, provided field Hk2 , at which point the resistance was restored. On they were pure and free from strain, elemental superconduct- reducing the field some hysteresis was observed, with a ors exhibited complete flux exclusion, a reversible transition small amount of flux remaining in the sample at zero field.23 at a well-defined critical field, and a final state independent Thus was type-II superconductivity recognized. It also ap- of the magnetization history. Alloys, on the other hand, peared that Silsbee’s hypothesis was obeyed by alloys, if the showed gradual flux penetration starting at a field below, and critical current was related to the lower critical field. finishing at a field somewhat higher, than the critical field Mendelssohn essayed an ingenious explanation for the typical of a pure element. In decreasing fields the magneti- two critical fields, the hysteresis and flux trapping, with his zation of alloys was hysteretic, and residual trapped flux was ‘‘sponge’’ model.24 This model postulated that a sponge or often retained when the applied field had returned to zero. three-dimensional network of superconductor with a high The so-called hard elemental superconductors such as Ta and critical field permeated the main body of the superconductor Nb showed behavior similar to that of alloys. with a lower critical field. Flux penetration would commence The research at Kharkov continued with careful magne- once the external field exceeded the critical value for the tization measurements on single and polycrystalline pure body of the superconductor, but penetration would not be metals, and on single alloy crystals of Pb–Bi, Pb–In, Pb–Tl, complete until the critical field of the sponge was reached. and Hg–Cd. Shubnikov’s final contribution to the critical On reducing the field, the meshes of the sponge would trap current story was systematic magnetization measurements on flux, accounting for hysteresis. The nature of the sponge was a series of PbTl single crystals of differing compositions.29 716 Low Temp. Phys. 27 (99–10), September–October 2001 D. Dew-Hughes

These showed that the change from ideal to alloy behavior point r is determined by the value of the magnetic vector occurred at a particular concentration of the alloying addi- potential A(r). In Pippard’s nonlocal modification of the tion. For lesser concentrations the alloy behaved as a pure London theory31 the current density at r is determined by A ␰ metal. As the concentration was increased above this particu- averaged over a volume of dimensions 0 . An electron trav- lar value, the field at which flux began to penetrate de- eling from a normal to a superconducting region cannot creased, and the field at which resistance was restored in- change its wave function abruptly; the change must take creased, with increasing concentration of alloying element. A place over some finite distance. This distance is called the ␰ clear picture of the change from what is now recognized as ‘‘range of coherence,’’ 0 . Pippard estimated that, for pure ͑ ͒ ␰ Ϸ ␮ type-I superconductor to type-II superconductor was pre- or clean metals, 0 1 m. The Pippard theory introduces sented, although there was an absence of cross-referencing modifications to the penetration depth. For a clean supercon- between the Ukrainian and Western European work. The the- ductor, clean in this case meaning that the normal electron ӷ␰ oretical explanation of the two types of superconductor was mean free path l 0 , the penetration depth is given by still missing, as was any understanding of what really deter- ␭ ϭ͓͑) ␲͒␰ ␭2 ͔1/3 ͑ ͒ mined the critical current density. No further progress had ϱ /2 0 L , 12 been made on these two problems when once again work on ␭ where L is the value of the penetration depth in the London superconductivity was frustrated by global conflict. theory. Ӷ␰ For alloy, or dirty superconductors, in which l 0 , the THEORETICAL ADVANCES 1945–1960 theory gives a new, much greater, value for the penetration depth, With the cessation of hostilities, renewed interest was ␭ϭ␭ ␰ ͒1/2 ͑ ͒ taken in superconductivity. Helium gas was now much more L͑ 0 /l , 13 readily available, its production having been accelerated by the needs of the US Navy for balloons. The development of and also a much reduced value for the coherence length, the Collins liquefier allowed many more physics laboratories ␰ ϭ ␰ ͒1/2 ͑ ͒ d ͑ 0l . 14 to indulge in studies at liquid-helium temperatures. However, the most startling advances were made on the theoretical The Ginsburg–Landau ␬ can be shown to be approximately front. equal to ␭/␰, and for a dirty superconductor with l very small, In 1950 Ginsburg and Landau, at the Institute for Physi- i.e., high electrical resistivity in the normal state, ␬ can be cal Problems in Moscow, published their phenomenological quite large, e.g., ϳ25 for niobium-based alloys and com- theory.30 They ascribed to the superconductor an order pa- pounds, and Ͼ100 for mixed oxide high-temperature super- rameter, ⌿, with some characteristics of a quantum- conductors. mechanical wave function. ⌿ is a function of temperature The next theoretical development was the formulation of and magnetic vector potential. The Gibbs function is ex- the Bardeen–Cooper–Schrieffer ͑BCS͒ microscopic theory 32 panded in even powers of ⌿ about the transition tempera- for superconductivity. This theory, for which the authors ture, as in Landau’s theory of phase transitions, and terms to received the 1972 Nobel Prize for Physics, is now the ac- describe the magnetic energy and kinetic energy and momen- cepted theory for conventional superconductors. In supercon- tum of the electrons are included in their expression for the ductors below the transition temperature, electrons close to Gibbs function of a superconductor in an external field. At an the Fermi surface condense into pairs ͑Cooper pairs͒. These external surface their theory reproduces the results of the pairs are the charge carriers in superconductivity, and their London theory. They introduced a new parameter, character- charge q is equal to twice the charge on a single electron. ␬ϭ&␭2 ␮ ប The value of the flux quantum ⌽ ϭh/2eϭ2.07 istic of a particular superconductor, q 0Hc / . The 0 Ϫ15 problem that they set out to solve, following the earlier ϫ10 Wb. The pairs form under an attractive interaction speculations of H. London, was that of the surface energy mediated by lattice phonons. An energy gap appears in the between superconducting and normal regions in the same excitation spectrum for electrons at the Fermi level. Electron metal. Their results showed quite clearly that, if ␬ were to pairs, lattice phonons, and energy gaps in superconductivity have a value greater than 1/&, then superconductivity could had been postulated previously, but Bardeen, Cooper, and persist up to fields in excess of the critical field, given by Schrieffer were the first to put all of these together in one ϭ ␬ & theoretical framework. The energy gap is related to the criti- H ( / )Hc . Ignoring the pre-War work on alloys, they stated that for no superconductor was ␬Ͼ0.1, and therefore cal temperature: this result was of no interest! 2⌬Ϸ3.5kT . ͑15͒ Pippard, with wartime experience of microwave tech- c niques, was now at Cambridge, engaged in measurements of This represents the energy required to break up the Cooper microwave surface resistance in metals and superconductors. pairs. It is possible to derive from this another estimate of the The anomalous skin effect in impure metals had been ex- maximum current density, the depairing current. The depair- plained by nonlocal effects. The behavior of an electron was ing current density is that at which the kinetic energy of the not influenced by the point value of the electric and magnetic superconducting carriers exceeds the binding energy of the fields but by the value averaged over a volume of dimensions Cooper pairs. It is then energetically favorable for the con- equal to the electron mean free path l. By analogy with the stituent electrons in a pair to separate and cease to be super- explanation for the anomalous skin effect in metals, Pippard conducting. The change in energy during scattering is maxi- suggested that a similar nonlocality was appropriate to super- mized when the momentum change is maximized. This conductors. In the London theory, the current density at a occurs when a carrier is scattered from one point on the Low Temp. Phys. 27 (99–10), September–October 2001 D. Dew-Hughes 717

͑ ͒ Fermi surface to a diametrically opposite one, in total rever- pounds NbN 15 K in 1941, V3Si(17 K) in 1951, ͑ϳ ͒ sal of direction. The carrier velocity is given by the sum of Nb3Sn(18 K) in 1954, and the ductile alloys Nb–Zr 11 K ϩ Ϫ the drift and Fermi velocities; vd v f becomes vd v f . The in 1953 and Nb–Ti ͑ϳ10 K͒ in 1961. All of these were resulting change in kinetic energy is type-II superconductors, with upper critical inductions well 1 1 in excess of previously known materials. Bc2 was about 12 T ␦E ϭ m͑v Ϫv ͒2Ϫ m͑v ϩv ͒2ϭϪ2mv v . ͑16͒ for Nb–Ti and 25 T for Nb Sn. The pioneers in this work n 2 d f 2 d f d f 3 were the groups at Westinghouse and Bell Telephone The breaking of a pair followed by scattering causes a Laboratories.2 Whereas the critical temperature and critical change in energy inductions were intrinsic properties of the superconductor, ␦ ϭ ⌬Ϫ ͑ ͒ the critical current density was found to be strongly depen- Es 2 2mvdv f . 17 dent upon the metallurgical state of the material. In two- ␦ For spontaneous depairing to occur, Es must be negative, phase alloys J was influenced by the size and dispersion of ⌬ c i.e., the drift velocity must be greater than /mv f . The de- the second phase particles.35 In the niobium-based ductile pairing current density Jd , which is just the drift velocity BCC alloys, it was found36 that cold deformation signifi- times the carrier density n and the carrier charge q, must cantly enhanced J . ϭ ⌬ c therefore be greater than Jd nq /mv f . When appropriate The problem of fabricating wire from the brittle interme- substitutions are made this expression for Jd can be shown to tallic Nb Sn was solved by filling niobium tubes with a mix- ␭ 3 reduce to Hc / , the previously quoted expression for the ture of Nb and Sn powders in the appropriate proportions, absolute maximum current density. Values of the depairing drawing to a fine wire, and reacting to form the compound.37 12Ϫ 13 2 current density lie in the range 10 10 A/m . This material had a current density of 109 A/m2 in an induc- Abrikosov, working in the same institute as Landau, 33 tion of 8.8 T. Similar wire was wound into a solenoid which made the fourth theoretical breakthrough in 1957. He pro- generated an induction of 2.85 T.38 If the reaction to form the duced a mathematical solution of the Ginsburg–Landau compound took place at the surface of the Nb and Sn par- ␬Ͼ & equations for the case when 1/ . His solution showed ticles, the compound could have formed as a three- that in a rising externally applied magnetic field, flux is ex- dimensional network, just as envisaged by Mendelssohn for cluded until a lower critical field, Hc1 , is exceeded. Above his sponge. Could the two-phase microstructure in the lead Hc1 flux penetrates in the form of flux vortices, or flux lines, alloys, or the dislocations introduced by deformation of the each carrying a quantum of flux ⌽ , directed parallel to the 0 ductile transition metal alloys,39 constitute the elements of field. The structure of these flux vortices is a normal core, of Mendelssohn’s sponge? Or was there an alternative scenario? radius ␰, containing the flux that is supported by supercur- If the flux vortices in the mixed state were able to interact in rents circulating over a radius ␭. As the applied field is in- some way with the microstructure, this interaction could im- creased, more flux penetrates until the density of the flux pede both the ingress of flux in a rising field and the egress lines is such that the normal cores begin to overlap. This of flux in a falling field. This would lead to the magnetic occurs at the upper critical field, H ϭ&␬H c2 c0 hysteresis observed in the materials. Flux gradients resulting ϭ⌽ /2␲␮ ␰2. The regime between the lower and upper 0 0 from nonuniform distributions of vortices can be equated critical fields is known as the ‘‘mixed state.’’ The mutual with currents. repulsion between the flux vortices, in the absence of any A current flowing in a superconductor in the mixed state other forces acting upon them, results in the formation of a will exert a Lorentz force on the flux vortices, F ϭJϫB triangular flux line lattice ͑FLL͒. The parameter of this lattice L(V) per unit volume of superconductor, or F ϭJϫ⌽ per unit is a ϭ1.07(⌽ /B)1/2, where B is the local value of the mag- L(l) 0 0 0 length of vortex, where ⌽ is a vector of strength ͉⌽ ͉ di- netic induction in the superconductor. Despite being pub- 0 0 rected along the vortex. The force acts in a direction normal lished in translation, Abrikosov’s paper took some time to be to both flux and current. Unless otherwise prevented, the fully appreciated in the West. vortices will move in the direction of this force, and in so In 1960 Gor’kov derived the constants in the Ginsburg– doing induce an electric field EϭvϫB, where v is the ve- Landau theory from the BCS theory.34 This trilogy of Rus- locity of the vortices. The superconductor now shows an sian theoretical work is collectively referred to as the GLAG ͑ ͒ induced resistance, the value of which approaches that of the Ginsburg–Landau–Abrikosov–Gor’kov theory. Supercon- ␳ ␬Ͼ & normal state, n , as the magnetic induction rises to Bc2 , the ductors with values of 1/ , which exhibit the mixed 40 state, are known as type-II superconductors. For any super- upper critical induction. The critical current is that current conductor, as the normal state mean free path of the elec- at which a detectable voltage is produced across the super- trons, l, is reduced, ␰ gets smaller, ␭ gets larger, and ␬ in- conductor, and is therefore that current which just causes the creases. Alloying, by reducing l, raises ␬. This explains vortices to move. If there is no hindrance to the motion of the Shubnikov’s observation that the change from type-I to vortices, then above Bc1 the critical current is zero and the type-II behavior, or the onset of the Shubnikov phase, occurs magnetization is reversible. The moving vortices do experi- at a particular alloy concentration.29 ence a viscous drag, originating from dissipation in the nor- ϭ␩ mal cores. This force Fv v, where the coefficient of vis- cosity ␩ϭ⌽ B/␳ ͑Ref. 40͒. If the vortices interact with APPLIED SUPERCONDUCTIVITY 1960–1986 0• n microstructural features in the body of the superconductor, The experimentalists had not been idle during this pe- such as impurities, crystal defects, and second-phase precipi- riod. New superconductors, showing a steady increase in tates, they can be prevented from moving and become critical temperature, had been discovered: the brittle com- pinned. The pinning force Fp is a function of the microstruc- 718 Low Temp. Phys. 27 (99–10), September–October 2001 D. Dew-Hughes ture and the local value of the induction. If the current den- pirical relations have been used to describe the dependence sity is such that the Lorentz force is less than the pinning of critical current on local magnetic induction. Surprisingly force, no movement of vortices will occur, and no voltage the simplest possible relation, which, as it turns out, fits the will be detected in the superconductor. If the current is in- data for commercial Nb–Ti conductor, namely Jc(B) ϭ Ϫ ϭ creased to a value at which the Lorentz force exceeds the Jc(0)(1 b), where b B/Bc2 is the reduced induction, pinning force, vortices will move and a voltage will be de- has been ignored. tected. The critical current density is that value of the current The problem of calculating critical currents from known density at which the vortices will begin to move, i.e., when details of the microstructure bears some relation to that of ϭ ϭ FL Fp ; thus giving Jc F p /B. calculating the mechanical properties of a structural alloy, or The sponge hypothesis was tested by studying an artifi- the magnetization curve of a magnetic material. In the case cial sponge fabricated by impregnating porous borosilicate of structural alloys, elastic inhomogeneities impede the glass with pure metal superconductors.41 The pores were in- movement of crystal dislocations. In the case of magnetic terconnected and had a diameter of 3–10 nm. At the Inter- materials, inhomogeneities in the magnetic properties im- national Conference on the Science of Superconductivity, pede the motion of domain walls. In superconductors the held at Colgate University the following year, the majority presence of inhomogeneity in the superconducting properties opinion swung in favor of pinning of flux vortices as the will impede the motion of flux vortices, and superconductors origin of magnetic hysteresis and the determinant of critical with strong pinning have been referred to as hard supercon- currents.42 Nevertheless, Bean’s experiments on the artificial ductors. The relation between microstructure, the properties sponge were important in leading to the concept of the criti- of the vortex lattice, and critical currents has been the subject cal state. Bean analyzed his results of magnetization mea- of several reviews, the most notable of which is that of surements on the assumption that each filament of the sponge Campbell and Evetts.45 carried either its critical current, or no current at all. As the Three factors must be considered in calculating pinning external field is raised, currents are induced in the outer fila- forces: the nature of the microstructural features, or pinning ments, shielding the inner filaments from the field. The field centers, responsible for pinning; the size, dispersion, and to- is able to penetrate only when the outer filament current at- pography of these pinning centers; and the rigidity of the tains its critical value. Filaments progressively carry the criti- flux-line lattice. The nature of the pinning center determines cal current until the flux has penetrated to the center of the the physical basis for the pinning force. A ferromagnetic pre- sample. Reducing the field to zero leaves current flowing in cipitate will react very strongly with a flux line.46 In most all the filaments, and flux is trapped in the sample. Applying cases the pins are either nonsuperconducting precipitates or a field in the opposite direction causes a progressive reversal voids,47 or regions whose superconductivity is modified, of the critical current in the filaments. Bean assumed a criti- such as dislocation tangles, grain and subgrain boundaries. cal current in the filaments independent of field. The model By passing through these regions the flux vortices reduce can be modified to include a field dependence of the critical their length, and hence their energy, in the superconductor. current, leading to a more realistic hysteresis curve. The size of pins is important, since if they have a dimension The notion that the current in a superconductor is either significantly less than the coherence length ␰, their effective- everywhere equal to the critical current or zero transfers ness is reduced by the proximity effect.48 If they have dimen- readily to the concept of a pinned Abrikosov vortex lattice. sions of the order of the penetration depth ␭, then local mag- In terms of magnetization, as the external field is raised, netic equilibrium within the pin can be established, vortices move into the superconductor. Their motion is re- magnetization currents will circulate around the pin, and the sisted by the pinning forces, and local equilibrium is estab- vortices will interact with these currents.49 The number of lished. At each point on the invading flux front the Lorentz pin–vortex interactions is determined by the dispersion of force exactly balances the pinning force, and the local cur- the pins. The topography decides whether the vortices, once rent density is equal to the local value of the critical current unpinned, must cut across the pins or are able to slide round density. The superconductor is in the critical state,43 a term them. borrowed from soil mechanics. A heap of soil or sand, or The lattice rigidity is important as, if the pinning centers snow on an alpine hillside, will come to equilibrium with a are randomly distributed, a rigid lattice will not be pinned. In slope of gradient determined by gravity and friction. The practice the lattice is not rigid, and three responses to the addition of more material to the pile will cause a slide until pinning or Lorentz forces imposed upon it can be recog- equilibrium is re-established. The slope is metastable, and nized. These forces may be such as to cause local elastic any disturbance will result in an avalanche. A similar situa- distortion of the lattice; they may exceed the yield strength tion obtains in a superconductor in the critical state. Any of the lattice, causing local plastic deformation; or they may force acting so as to try to move a flux vortex is just opposed exceed the shear strength of the lattice. Whichever of these by an equal and opposite pinning force. An imposed distur- possibilities actually occurs provides the answer to what is bance, resulting from either a change in the external mag- known as the summation problem. If the lattice undergoes netic field or in a transport current, leads to a redistribution elastic distortion, the situation involves collective pinning.50 of flux until the critical state is restored. Spectacular The vortices are weakly pinned and the supercurrent densi- flux avalanches, or jumps, have been observed in ties are too low to be of practical interest. This situation will superconductors.44 The one difference in the superconductor not be considered further. If the pinning forces are such as to is that, as the pinning force is a function of the local induc- cause local plastic deformation of the vortex lattice, the vor- tion, the slope of the flux front is not constant. Several em- tices will position themselves so as to maximize the pinning Low Temp. Phys. 27 (99–10), September–October 2001 D. Dew-Hughes 719

␰ interaction. Each vortex can be assumed to act individually, interaction distance is . Taking Nb3Sn as an example, with ϭ ␰ϭ ϫ Ϫ9 and the global pinning force is just the direct sum of the Bc 1 T and 3.6 10 m, and considering that b(1 individual forces. If the pinning forces are greater than the Ϫb) has a maximum value of 0.25 at bϭ0.5, i.e., at B shear strength of the vortex lattice, some vortices may re- ϭ12.5 T, we find main pinned, while the main part of the lattice shears past them.51 However, this can only happen if there are paths b͑1Ϫb͒ 5.5ϫ1013 0.25 J ϭ ϭ • down which the vortices can move without traversing any c 4.4␲ϫ10Ϫ7 3.6ϫ10Ϫ9B 12.5 pins.52 • ϫ ϭ 12 2 ͑ ͒ If the experimental critical Lorentz force Jc B, deter- 10 A/m . 21 mined from transport current measurements, is plotted versus the reduced value of the applied magnetic induction, b,itis Thus the maximum possible critical current density due to found that, for a given sample, results at different tempera- pinning is about one-tenth of the depairing current density. In tures lie on one master curve.53 The master curve takes the practice, of course, it is impossible to achieve this idealized form microstructure; maximum critical current densities due to pinning are about one-hundredth of the above estimate. ϭ pϩq p͑ Ϫ ͒q ͑ ͒ JcB const•Bc2 b 1 b , 18 The two conventional superconductors in commercial where the temperature dependence is incorporated in the production, the ductile transition metal alloy Nb–Ti, and the temperature dependence of the upper critical induction. This intermetallic compound Nb3Sn, will now be examined in the is known as a scaling law. The values of the exponents p and light of the ideas expressed in the previous paragraphs. In order to confer stability, these conductors are fabricated as q are peculiar to the particular pinning mechanism. Scaling 55 laws are fundamental to flux pinning.54 As an example, pin- many fine filaments of superconductor in a copper matrix. ning by nonsuperconducting precipitates will be considered. In the case of Nb–Ti, rods of the alloy are inserted in a If an isolated vortex intersects a spherical particle of normal copper matrix, and drawn down, often with repeated bun- material of diameter D, a volume of vortex core D␲␰2 is dling, drawing, and annealing schedules, to produce a mul- removed from the system. Associated with the vortex core is tifilamentary composite wire. Extensive transmission elec- 2 ␮ tron microscope studies on pure Nb and V, and alloys of an energy per unit volume Bc /2 0 . Thus the energy of the ␲␰2 2 ␮ Nb–Ta, Nb–Zr, Nb–Ti, and Mo–Re, after cold deformation system is lowered by an amount D Bc /2 0 . The force to move the vortex from a position in which it passes through and annealing, have shown conclusively that, in these ductile the center of the particle, to a position outside the particle, is metals, pinning is due to an interaction between flux lines and tangles of dislocations or cell walls, and not individual this change in energy divided by an interaction distance, 56 which in this case is clearly the diameter of the particle. Thus dislocations. In these tangles the normal electron mean free the force to depin an isolated vortex from a normal particle is path will be less than its value in the dislocation-free regions, ␬ ␲␰2B2/2␮ . The total pinning force per unit volume is the and the local value of will be increased. This led to the idea c 0 of ⌬K pinning,57,58 the theory for which was developed by single pin force multiplied by the number of active pins per 59 unit volume. In this case this latter quantity is approximately Hampshire and Taylor. The superconducting filaments in equal to the total length of vortices per unit volume, B/⌽ , Nb–Ti have a heavily deformed microstructure, with grains, 0 subgrains, and nonsuperconducting ␣-Ti particles elongated multiplied by the volume fraction of particles, V f . There is an additional effect to be taken into account. In the flux-line in the direction of drawing. The current flow is parallel to lattice, of reduced induction b, the density of superelectrons, this elongated microstructure, and the Lorentz force acting and hence the superconducting condensation energy, is re- on the flux vortices is such as to drive them across the sub- duced by a factor (1Ϫb).45 The pinning force per unit vol- grain and normal particle boundaries. Pinning occurs at these ⌬ ume is thus boundaries and is a mixture of normal-particle and K pin- ning, with a pinning function in which the critical Lorentz B2 B force J B is proportional to b(1Ϫb).54 The critical current is ϭ␲␰2 c ͑ Ϫ ͒ ͑ ͒ c JcB ␮ 1 b ⌽ V f . 19 associated with the unpinning of flux vortices from these 2 0 0 boundaries. The derivation of the pinning function is along ϭ⌽ ␲␰2 With the use of the expression for Bc2 0/2 , this be- similar lines to that described above for normal particles. comes Theory and experiment are well matched.52 The above ex- 2 pression seems to hold whenever the critical current is deter- Bc J Bϭ b͑1Ϫb͒V . ͑20͒ mined by flux pinning with a density of pins less than the c 4␮ f 0 density of flux lines. The b term arises because, as the den- The above derivation assumes only one vortex is pinned at sity of flux lines increases, so does the total length of line each particle, and therefore the particle size must be less than pinned. The (1Ϫb) term represents the decrease in super- the intervortex spacing. Based on the above expression, it is conducting order parameter with increasing induction. possible to make an estimate of the maximum pinning force, The other commercial conductor is based on the inter- and hence the maximum current density. In order to maxi- metallic A15-type compound Nb3Sn. Multifilamentary con- mize the pinning force, all vortices must be pinned over their ductor is fabricated by some variant of the bronze process. In entire length. This would require a microstructure consisting the original version of this process, rods of niobium are in- of continuous rods of nonsuperconductor, with diameter ϳ␰, serted in a copper/tin bronze ingot as matrix, and drawn, parallel to the applied field, and at a spacing equal to that of again with rebundling, to form a composite of fine niobium the vortex lattice. In this case V f is effectively 1, and the filaments in the bronze matrix. Reaction between the tin con- 720 Low Temp. Phys. 27 (99–10), September–October 2001 D. Dew-Hughes tent of the bronze and the niobium at an elevated temperature tions of 40 T.62 As the temperature is increased, all regions of converts the latter into Nb3Sn filaments. This procedure is the curve move to lower values of field and critical current necessary, as the intermetallic compound is brittle and non- density. In particular, the cutoff field decreases and the deformable. The critical Lorentz force in these materials is ͑negative͒ slope of the middle region increases. The signifi- found to obey a scaling law similar to that postulated by cant fundamental differences between low-temperature and Kramer,51 namely b1/2(1Ϫb2). The critical current density high-temperature superconductors are that the latter are an- increases as the grain size decreases, as would be expected if isotropic and have rather small coherence lengths. Structur- the pinning occurred at the grain boundaries, and as it does ally the mixed oxide superconductors are tetragonal, or in Nb–Ti. The (1Ϫb2) term has been taken to be indicative nearly-tetragonal, with lattice parameters a and b lying in the of some flux shearing process, as the C66 modulus of the range 0.375–0.395 nm and the c-axis parameter 3–12 times 2 flux-line lattice varies as (1Ϫb ) at high values of b.Itisnot greater. This structural anisotropy leads to anisotropy in the immediately obvious as to why these two types of material physical properties of the compounds. In single crystals, the should behave in such different fashion, as their supercon- critical current density in the ab plane is many times greater ducting parameters and scale of microstructure are not vastly than that in the c direction, normal to the ab plane. The different. However examination of the microstructure of superconducting coherence length ␰ is small in these com- Nb3Sn reveals it to be very different from that of Nb–Ti. pounds; that in the c direction is just a few tenths of a na- This is not at all unexpected, due to the very different ways nometer in length, of similar magnitude to the region of crys- in which two microstructure are generated. That of bronze- tallographic disturbance in the boundary between two grains. processed Nb3Sn consists of columnar grains whose axes are The consequence of this small range of coherence is that 60 normal to the axes of the filaments. The Lorentz force will grain boundaries in high-temperature superconductors act as act parallel to some of these boundaries, driving the flux weak links, i.e., the superconducting wave functions in adja- lines along them rather than across them. A path is thus pro- cent grains are only weakly coupled to one another. The vided down which flux can shear, and the author has put overall critical transport current density in a superconductor forward a mechanism of flux-lattice dislocation-assisted 52 is determined by whichever is the lesser of the intragrain or shear. Values of the critical Lorentz force predicted on this the intergrain current densities. The intragrain current den- model are close both to the Kramer law and to observation; sity is controlled by flux pinning, the intergrain current den- in addition the model predicts an inverse dependence of J c sity is a measure of the ability of current to flow from one on grain size, as is observed experimentally but not predicted grain to an adjacent grain. This latter depends upon the on the Kramer theory. An alternative approach treats flux strength of the superconducting link across the boundary, and pinned at grain boundaries as Josephson vortices.61 Trans- in the case of anisotropic superconductors, upon the relative verse unpinning, with vortices crossing grain boundaries as orientation between the two grains.63 The initial rapid drop in in Nb–Ti, leads to the b(1Ϫb) scaling law, while longitudi- J with field is due to many weak links between grains being nal unpinning, with vortices traveling along grain boundaries c progressively switched off as the field is increased.64 as proposed for Nb Sn, leads to the b1/2(1Ϫb2) scaling law. 3 The current that is left is now being carried by the few strong links that exist between the grains, and the number of HIGH-TEMPERATURE SUPERCONDUCTORS these is relatively insensitive to magnetic field. The strength The immediate expectation from the discovery of the of supercurrent depends upon the proportion of grain bound- high-temperature, mixed copper oxide superconductors was aries that are strong links. Many models have been proposed that these materials could be exploited at 77 K to build elec- to account for the manner in which current is transferred tromagnets that would compete with permanent magnets, of- from grain to grain in anisotropic mixed oxide 65 fering inductions in excess of 2 T. At low temperatures, the superconductors. The conclusions from these models, con- high critical fields would allow of competition with low- firmed by experience, is that the proportion of strong links temperature superconductors, and the 21 T maximum induc- between grains, and hence the intergrain current, is maxi- tion available from existing A15 conductor would be mized by grain alignment. The material is textured so that exceeded. These high hopes have met with disappointment; the c axis of the grains is close to being normal to the direc- the critical current densities, especially in high magnetic tion of current flow, and that the ab planes of the grains are fields, are much less than those in low-temperature supercon- in near parallelism to one another. In effect, the conductor ductors. must be as close to being a single crystal as possible. Typically, the critical current density as a function of Once a degree of texture has been established, the cur- applied induction for a high-temperature superconductor rent density is further determined by flux pinning. A fully shows three regimes: an initial region in which the critical textured material will carry no appreciable current density if current decreases rapidly as soon as the field is turned on; a the pinning is weak. Conversely, a material with strong pin- region, which can be linear, falling slowly with increasing ning will also have a low critical current density if there is no field, and a third region in which the critical current falls to texture. In anisotropic materials the pinning of flux is also zero. The middle region may appear to be perfectly horizon- anisotropic.66 The pinning strength is a function of the direc- tal, indicating no dependence of critical current on applied tion of an external magnetic field relative to the ab planes of field. It may also extend to very high fields, especially in the superconductor. The critical current density is much Bi-2212 at temperatures below 20 K. An extreme example is higher with the field parallel to the ab planes than when it is a sample of spray-pyrolized TI-1223, in which the critical perpendicular to them. The high-temperature superconduct- current density at 4.2 K is constant with field up to induc- ing compounds consist of groups of one, two, or three copper Low Temp. Phys. 27 (99–10), September–October 2001 D. Dew-Hughes 721 oxide layers, which are responsible for the superconductivity, dismissed the possibility of superconductors having values of separated by layers of other oxides that are essentially insu- ␬Ͼ1/&. Abrikosov’s ideas were slow to be appreciated. lating. With the field lying parallel to the ab planes, the vor- One is tempted to ask, ‘‘Would the first proper applications tices will tend to place themselves in the insulating layers. of superconductivity, in high field magnets, have arisen ear- The pinning mechanism, known as intrinsic pinning, is simi- lier than ϳ1960 if these delays had not occurred?’’ The an- lar to that by normal particles as discussed above for low- swer is almost certainly ‘‘no.’’ The applications were condi- temperature superconductors. The maximum critical current tional upon the discovery and development of materials with density should be of the same magnitude as that estimated in the ability to carry high currents in high magnetic fields. Eq. ͑20͒. The density of pins is much greater than the density These discoveries did not rely upon any phenomenological of flux lines, explaining the relative insensitivity of the cur- or theoretical developments, but were, as are so many useful rent density to external magnetic field in the middle region of discoveries, purely empirical. the Jc versus B curve. When the applied field is normal to the The critical current density in both low-temperature and ab planes, the intrinsic pinning no longer acts to hinder flux- high-temperature superconductors is controlled by their mi- line motion; the critical current densities are much lower crostructure. Flux pinning in the ductile alloys based on nio- than when the field is parallel to the planes. The situation is bium occurs at dislocation tangles, subgrain boundaries, and made worse by the fact that flux lines normal to the ab interfaces with nonsuperconducting second phases ␣-Ti. Flux planes tend to split into ‘‘pancakes.’’ 67 This tendency is shear along columnar grain boundaries seems to be the con- greater the greater the ratio of nonsuperconducting oxide trolling mechanism in the bronze-route A15 materials. In the layer thickness to superconducting oxide layer thickness, and high-temperature superconductors microstructural control hence the degree of anisotropy in the material. The anisot- must provide both a high degree of texture and flux pinning. ropy can be reduced, and flux pinning can be enhanced, by The next challenge will be to control the microstructure of chemical substitution that distorts the crystal structure, by the MgB2. addition of nonsuperconducting phases, and by irradiation. For the pre-War history of superconductivity, I have As the applied field continues to increase, a value is drawn heavily upon P. Dahl’s book Superconductivity.71 reached at which the critical current falls to zero. This is the irreversibility field, above which it becomes impossible to *E-mail: [email protected] pin flux. Irreversibility in magnetization experiments also disappears. The magnitude of the irreversibility field de- creases as the anisotropy and tendency to form pancake vor- 1 H. K. Onnes, Supplements to the Communications of the Physical Labo- tices increases. There is controversy as to the origin of the ratory of Leiden University 34b,55͑1913͒. 2 ͑ ͒ irreversibility field. Arguments persist as to whether it is J. K. Hulm, J. E. Kunzler, and B. T. Matthias, Phys. Today 34 1 ,34 ͑1981͒. caused by flux-lattice melting or by thermally activated de- 3 H. K. Onnes, Communications of the Physical Laboratory of Leiden Uni- pinning. What is interesting is that the critical Lorentz force versity 133a,3͑1913͒. in high-temperature superconductors in many cases follows 4 H. K. Onnes, ibid. 133b,29͑1913͒. 5 ͑ ͒ scaling laws similar to those found for low-temperature su- H. K. Onnes, ibid. 133d,51 1913 . 6 H. K. Onnes, ibid. 139f,66͑1914͒. perconductors. The one difference is that the reduced induc- 7 F. B. Silsbee, Journal of the Washington Academy of Sciences 6,79 tion used in the scaling laws is that relative to the irrevers- ͑1916͒. ibility field rather than the upper critical field. There are 8 F. B. Silsbee, Proc. Natl. Acad. Sci., USA 13,516͑1927͒. 9 many examples of this in the literature. Scaling with the W. Tuyn and H. K. Onnes, Communications of the Physical Laboratory, University of Leiden, 174a,21͑1926͒. irreversibility field indicates that this field is an intrinsic 10 W. Tuyn and H. K. Onnes, J. Franklin Inst. 201, 379 ͑1926͒. property of the flux-line lattice. 11 W. J. De Haas, G. J. Sizoo, and H. K. Onnes, Physica 5, 447 ͑1925͒. 12 G. J. Sizoo, W. J. De Haas, and H. K. Onnes, in Proceedings of the The recently discovered superconductor MgB2, with a 68 Amsterdam Akademie der Wetenschappen 29, 221 ͑1926͒. critical temperature of 39 K, appears to be a conventional 13 W. J. De Haas and J. Voogd, ibid. 32,874͑1929͒; 33,262͑1930͒. low-temperature superconductor, with well-coupled grains 14 W. H. Keesom, Physica 2,36͑1935͒. and strong bulk pinning.69 Transport current densities of 15 W. J. De Haas and J. Voogd, in Proceedings of the Amsterdam Akademie 8 2 der Wetenschappen 34,63͑1931͒. 10 A/m , measured in self-field at 4.2 K, have been reported 16 70 M. Von Laue, Phys. Z. 33, 793 ͑1932͒. in wires fabricated from this material. The nature of the 17 W. Meissner and R. Ochsenfeld, Die Naturwissenschaften 21,787͑1933͒. pinning sites has not yet been determined. 18 J. N. Rjabinin and L. V. Shubnikov, Nature ͑London͒ 134,286͑1934͒. 19 C. J. Gorter and H. B. J. Casimir, Phys. Z. 35, 963 ͑1934͒. 20 F. London and H. London, Proc. R. Soc. London, Ser. A 149,71͑1935͒. CONCLUSIONS 21 T. C. Keely, K. Mendelssohn, and J. R. Moore, Nature ͑London͒ 134,773 ͑1934͒. The history of the experimental facts, and the theories 22 W. J. De Haas and J. M. Casimir-Jonker, ibid. 135,30͑1935͒. developed therefrom, that have defined the understanding of 23 J. N. Rjabinin and L. V. Schubnikov, Phys. Z. Sowjetunion 7, 122 ͑1935͒. the factors that control critical currents in both low- and 24 K. Mendelssohn and J. R. Moore, Nature ͑London͒ 135,826͑1935͒. 25 ͑ ͒ high-temperature superconductors, has been delineated. At C. J. Gorter, Physica 2,449 1935 . 26 B. B. Goodman, IBM J. Res. Dev. 6,63͑1962͒. several critical stages opportunities have been missed. Onnes 27 H. London, Proc. R. Soc. London, Ser. A 152, 650 ͑1935͒. failed to connect critical fields with critical currents. The 28 A. Shalnikov, Nature ͑London͒ 142,74͑1938͒. Meissner effect was discovered rather later than it ought to 29 L. V. Schubnikov et al., Phys. Z. Sowjetunion 10,165͑1936͒. 30 ´ ͑ ͒ have been. In the late 1930s there was a lack of cooperation V. L. Ginzburg and L. D. Landau, Zh. Eksp. Teor. Fiz. 20, 1064 1950 . 31 A. B. Pippard, Proc. R. Soc. London, Ser. A 216,547͑1953͒. between the Leiden and Oxford groups on the one hand, and 32 J. Bardeen, L. N. Cooper, and J. R. Schrieffer, Phys. Rev. 108, 1175 the Kharkov group on the other hand. Ginsburg and Landau ͑1957͒. 722 Low Temp. Phys. 27 (99–10), September–October 2001 D. Dew-Hughes

33 A. A. Abrikosov, Zh. E´ ksp. Teor. Fiz. 5, 1442 ͑1957͓͒Sov. Phys. JETP 5, 52 D. Dew-Hughes, Philos. Mag. B 55, 459 ͑1987͒. 1174 ͑1957͔͒. 53 W. A. Fietz and W. W. Webb, Phys. Rev. 178, 657 ͑1969͒. 34 L. P. Gor’kov, ibid. 10,998͑1960͒. 54 D. Dew-Hughes, Philos. Mag. 30,293͑1974͒. 35 J. D. Livingston, Phys. Rev. 129, 1943 ͑1963͒; Acta Metall. 11, 1371 55 M. N. Wilson, C. R. Walters, J. D. Lewin, and P. F. Smith, J. Phys. D 3, ͑1963͒; J. Appl. Phys. 34,3028͑1963͒; Rev. Mod. Phys. 36,54͑1964͒. 1518 ͑1970͒. 36 M. S. Walker, R. Stickler, and F. E. Werner, in Metallurgy of Advanced 56 A. V. Narlikar and D. Dew-Hughes, Phys. Status Solidi 6,383͑1964͒; Electronic Materials, edited by G. E. Brock ͓Interscience, New York J. Mater. Sci. 1, 317 ͑1966͒. ͑1963͒,p.49͔. 57 D. Dew-Hughes, Mater. Sci. Eng. 1,2͑1966͒. 37 J. E. Kunzler, E. Buehler, F. S. L. Hsu, and J. E. Wernick, Phys. Rev. Lett. 58 D. Dew-Hughes and M. J. Witcomb, Philos. Mag. 26,73͑1972͒. ͑ ͒ 6,89 1961 . 59 R. G. Hampshire and M. T. Taylor, J. Phys. F: Met. Phys. 2,89͑1972͒. 38 ͑ ͒ L. C. Salter et al.,inHigh Magnetic Fields,H.H.Kolmet al. Eds. , John 60 C. S. Pande, Metallurgy of Superconducting Materials, edited by T. S. , New York ͑1962͒,p.344. ͑ ͒ 39 Luhman and D. Dew-Hughes, Academic Press, New York 1979 ,p.171. R. Schaw and D. E. Mapother, Phys. Rev. 118, 1474 ͑1960͒. 61 ͑ ͒ 40 J. McDonald and E. Barzi, IEEE Trans. Appl. Supercond. 11,3884 2001 . Y. B. Kim, C. F. Hempstead, and A. R. Strnad, Phys. Rev. 139,A1163 62 D. T. Ryan et al., IEEE Trans. Magn. 32,2803͑1996͒. ͑1965͒. 63 D. Dimos et al., Phys. Rev. Lett. 61,219͑1988͒. 41 C. P. Bean, M. V. Doyle, and A. G. Pincus, Phys. Rev. Lett. 9,93͑1962͒. 64 R. L. Peterson and J. W. Ekin, Phys. Rev. B 42, 8014 ͑1990͒. 42 See Rev. Mod. Phys. 36, ͑1964͒. 65 L. N. Bulaevskii et al., Phys. Rev. B 45, 2545 ͑1992͒; B. Hensel et al., 43 C. P. Bean, Phys. Rev. Lett. 8, 250 ͑1962͒; Y. B. Kim, C. F. Hempstead, Physica C 205, 329 ͑1993͒. and A. R. Strnad, Phys. Rev. 129, 528 ͑1963͒. 66 P. H. Kes et al., Phys. Rev. Lett. 64,1063͑1990͒. 44 J. E. Evetts, A. M. Campbell, and D. Dew-Hughes, Philos. Mag. 10, 339 67 J. R. Clem, Phys. Rev. B 43, 7837 ͑1991͒. ͑1964͒. 68 J. Nagamatsu et al., Nature ͑London͒ 410,63͑2001͒. 45 A. M. Campbell and J. E. Evetts, Adv. Phys. 21,199͑1972͒. 69 ͑ ͒ 46 ͑ ͒ M. Kambara et al., Semicond. Sci. Technol. 14,L5 2001 . T. H. Alden and J. D. Livingston, J. Appl. Phys. 37, 3551 1966 . 70 47 ͑ ͒ B. A. Glowacki et al., ibid. 14, 193 ͑2001͒. E. J. Kramer and H. L. Freyhardt, J. Appl. Phys. 51, 4930 1980 . 71 48 J. Friedel, P. G. de Gennes, and J. Matricon, Appl. Phys. Lett. 2,119 P. F. Dahl, Superconductivity, its Historical Roots and Development from ͑1963͒. Mercury to the Ceramic Oxides, American Institute of Physics, New York ͑ ͒ 49 A. M. Campbell, J. E. Evetts, and D. Dew-Hughes, Philos. Mag. 18, 313 1992 . ͑1968͒. 50 A. I. Larkin and Yu. N. Ovchinnikov, J. Low Temp. Phys. 34, 409 ͑1979͒. This article was published in English in the original Russian journal. Repro- 51 E. J. Kramer, J. Appl. Phys. 44, 1360 ͑1973͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Geometric edge barrier in the Shubnikov phase of type-II superconductors E. H. Brandt*

Max-Planck-Institut fu¨r Metallforschung, D-70506 Stuttgart, Germany; Institute for Superconducting and Electronic Materials, University of Wollongong, NSW 2522 Australia ͑Submitted April 25, 2001͒ Fiz. Nizk. Temp. 27, 980–990 ͑September–October 2001͒ In type-II superconductors the magnetic response can be irreversible for two different reasons: vortex pinning and barriers to flux penetration. Even without bulk pinning and in the absence of a microscopic Bean–Livingston surface barrier for vortex penetration, superconductors of nonellipsoidal shape can exhibit a large geometric barrier for flux penetration. This edge barrier and the resulting irreversible magnetization loops and flux-density profiles are computed from continuum electrodynamics for superconductor strips and disks of constant thickness, both without and with bulk pinning. Expressions are given for the field of first flux entry Hen and for the reversibility field Hrev above which the pin-free magnetization becomes reversible. Both fields are proportional to the lower critical field Hc1 but otherwise depend only on the specimen shape. These results for rectangular cross section are compared with the well- known reversible magnetic behavior of ideal ellipsoids. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1401181͔

1. SHUBNIKOV PHASE WITH ABRIKOSOV’S FLUX-LINE loops that may become normal conducting when the circu- LATTICE lating current exceeds some critical value. The true explana- tion of partial flux penetration was given in a pioneering Many metals, alloys, and compounds become supercon- work by Alexei Abrikosov in 1957.13 Abrikosov, a student of ducting when they are cooled below a transition temperature Lev Landau in Moscow, discovered a periodic solution of the T . This critical temperature ranges from T Ͻ1 K for Al, c c phenomenological theory of superconductivity conceived a Zn, Ti, U, and W and T ϭ4.15 K for Hg ͑the first supercon- c few years earlier by Ginzburg and Landau.14 Abrikosov in- ductor discovered, in 1911͒,1 through T ϭ9.2 K for Nb ͑the c terpreted his solution as a lattice of parallel flux lines, now elemental metal with the highest T ͒ and T Ϸ23 K for c c also called flux tubes, fluxons, or Abrikosov vortex lines. Nb Ge ͑the highest value from 1973 to 1986; see the over- 3 These flux lines thread the specimen, each carrying a quan- view in Ref. 2͒, to the large T values of the high-T super- c c ␾ ϭ ϭ ϫ Ϫ15 2 conductors ͑HTSCs͒ discovered in 1986,3 e.g., tum of magnetic flux 0 h/2e 2.07 10 T•m . At the ͑ ␦Ӷ ͒ 4 Ϸ center of a flux line the superconducting order parameter YBa2Cu3O7Ϫ␦ YBCO, 1 , with Tc 92.5 K, and 5,6 ␺(r) ͑the complex Ginzburg–Landau ͑GL͒ function͒ van- Bi Sr Ca Cu O ϩ␦ ͑BSCCO͒, with T up to 120 K, and 2 2 2 3 10 c ␺ϭ Ϸ␰ on up to Tl Ba Sr Ca Cu O , with maximum T ϭ127 K,7 ishes. The line 0 is surrounded by a tube of radius , 2 2 2 2 3 10 c ͉␺͉ and some Hg compounds which under pressure have reached the vortex core, within which is suppressed from its su- Ϸ 8,9 perconducting value ͉␺͉ϭ1 that it attains in the Meissner Tc 164 K; the just recently discovered ‘‘simple’’ super- ϭ 10 state. The vortex core is surrounded by a circulating super- conductor MgB2 has Tc 39 K. The superconducting state is characterized by the van- current J(r) which generates the magnetic field B(r) of the ishing of the electric resistivity ␳(T) of the material and by flux line. In bulk specimens the vortex current and field are ␭ the complete expulsion of magnetic flux, irrespective of confined to a flux tube of radius , the magnetic penetration depth; at large distances rӷ␭, the current and field of an whether the magnetic field Ba was applied before or after isolated vortex decay as exp(Ϫr/␭). cooling the superconductor below Tc . The existence of this Meissner effect proves that the superconducting state is a In thin films of thickness dӶ␭ the current and magnetic ␭ ϭ ␭2 thermodynamic state, which uniquely depends on the applied field of a vortex extend to the larger distance film 2 /d, field and temperature but not on previous history. As opposed the circulating current and the parallel magnetic field at large ӷ␭ 2 to this, an ideal conductor expels the magnetic flux of a distances r film decrease only as 1/r and the perpendicu- 3 suddenly switched on field Ba but also ‘‘freezes’’ in its inte- lar field as 1/r , and the vortex core has a wider radius Ϸ ␭ ␰2 1/3 ͑ ͒ rior the magnetic flux which was there before the conductiv- (12 film ) Refs. 15 and 16 . These thin-film results ity became ideal. have been applied to the high-Tc superconductors with lay- Lev Shubnikov realized that some superconductors do ered structure, defining the vortex lines as stacks of vortex not exhibit complete expulsion of flux, but the applied field disks ͑‘‘pancake vortices’’͒ in the superconducting CuO partly penetrates and the magnetization of the specimen de- layers.17 The coherence length ␰(T) and magnetic penetra- 11,12 ␭ pends on the magnetic history in a complicated way. tion depth (T) of the GL theory diverge at temperature Tc Ϫ Ϫ1/2 Early theories tried to explain this by a ‘‘spongelike’’ nature as (1 T/Tc) . of the material, which could trap flux in microscopic current The ratio ␬ϭ␭/␰ is the GL parameter of the supercon-

1063-777X/2001/27(9–10)/9/$20.00723 © 2001 American Institute of Physics 724 Low Temp. Phys. 27 (99–10), September–October 2001 E. H. Brandt

ϭ 2ϩ 2 1/2 ʈ ductor. Within GL theory, which was conceived for tempera- with r (x y ) and B z; K0(x) is a modified Bessel ␬ Ϫ Ӷ ␲ 1/2 tures close to the transition temperature Tc , is independent function with the limits ln(x)(x 1) and ( /2x) exp of T. Abrikosov’s flux-line lattice ͑FLL͒ exists only in mate- (Ϫx)(xӷ1). This field B(r) exactly minimizes the GL rials with ␬Ͼ1/&; these are called type-II superconductors free energy if the above variational ansatz ͉␺(r)͉2 is inserted. ␬Ͻ & as opposed to type-I superconductors, which have 1/ . The maximum field occurs in the vortex core, Bmax ϭ Ϸ ␾ ␲␭2 ␬Ϸ ͑ ␬2ӷ ͒ Type-I superconductors in a parallel applied field Ha B(0) ( 0/2 )ln 2Bc1 still for 2 1 . From this Ͻ Hc(T) are in the Meissner state, i.e., flux penetrates only B(r) one obtains the current density circulating in the ␭ Ͼ ϭ␮Ϫ1͉ Ј ͉ϭ ␾ ␲␭2␮ ␭ ␭ into a thin surface layer of depth (T), and at Ha Hc(T) vortex J(r) 0 B (r) ( 0/2 0)(r/ ˜r)K1(˜r/ ), 2 2 1/2 they become normal conducting. Here Hc(T) is the thermo- with ˜rϭ(r ϩ2␰ ) . Inserting for the modified Bessel Ϸ dynamic critical field. Type-II superconductors in a parallel function K1(x) the approximation K1(x) 1/x valid Ͻ р Ӷ applied field Ba Bc1(T) Bc(T) are in the Meissner state, for x 1, one obtains the maximum current density Jmax ϭ ϭ&␰ Ϸ␾ &␲␭2␰␮ ϭ 1/2 ϭ␾ i.e., no magnetic flux has penetrated, and their inner induc- J(r ) 0 /(4 0) (27/32) J0 where J0 0 / ϭ Ͻ Ͻ )␲␭2␰␮ tion is thus B 0; in the field range Hc1(T) Ha Hc2(T) (3 0) is the ‘‘depairing current density,’’ i.e., the the magnetic flux penetrates partly in the form of flux lines maximum supercurrent density which can flow within the ͑ Ͻ Ͻ␮ ͒ 24 Shubnikov phase or mixed state with 0 B 0Ha ; and at GL theory in planar geometry ͑see, e.g., Tinkham ͒. Thus, Ͼ у Ha Hc2(T) Hc(T) the material is in the normal conduct- for large ␬ӷ1 the field in the flux-line center is twice the ϭ␮ ing state, and thus B 0Ha . Hc1 and Hc2 are the lower and lower critical field, and the maximum vortex current is the upper critical fields. One has depairing current. A curious property of the flux-line lattice is its softness, ␾ Ϸ 0 ͑ ␬ϩ ͒ which is due to the long range interaction between the flux Hc1 ␲␭2␮ ln 0.5 , ␭ 4 0 lines over several penetration lengths , a distance which typically is much larger than the flux-line spacing. This leads ␾ ␾ to ‘‘nonlocal’’ elastic behavior and to highly dispersive elas- 0 0 H ϭ , H ϭ ϭ&␬H . tic moduli for compression ͓c (k)͔ and tilt ͓c (k)͔, while c &␲␰␭␮ c2 2␲␰2␮ c 11 44 2 0 0 ͓ Ӷ Ϸ Ϸ 2 ␮ the very small shear modulus c66 c11(0) c44(0) B / 0 for BϾ␮ H ͔ does not depend on the wave vector k of the All three critical fields vanish for T→T as T ϪT and have 0 c1 c c strain field.25 For more properties of the ideal and pinned an approximate temperature dependence ϰ1ϪT2/T2. c FLL, also in the highly anisotropic or layered high-T super- If the superconductor is not a long specimen in a parallel c conductors, see the reviews on Refs. 26 and 27, and for the field, then demagnetization effects come into play. For ellip- rather complex statistical theory of pinning and thermally soidal specimens with homogeneous magnetization the de- activated depinning of vortex lines and pancake vortices, see magnetizing field is taken into account by a demagnetization the review in Ref. 28. The properties of the ideally periodic factor N with 0ϽNϽ1. If NϾ0, flux penetration starts ear- FLL have recently been computed with high accuracy for the lier, namely, into type-II superconductors at HЈ ϭ(1 c1 entire ranges of the induction 0ϽBϽ␮ H and of the GL ϪN)H in the form of a FLL, and into type-I superconduct- 0 c2 c1 parameter 1/&Ͻ␬Ͻϱ by an iteration method.29 ors at HЈϭ(1ϪN)H in the form of normal conducting c c The present paper considers the magnetic behavior of lamellae; this ‘‘intermediate state’’ is described by Landau superconductors which are not long cylinders or ideal ellip- and Lifshitz;18 see also Refs. 19 and 20. GL theory yields soids but have a more realistic constant thickness, i.e., they that the wall energy between normal and superconducting have rectangular cross section in the planes containing the domains is positive ͑negative͒ for type-I ͑type-II͒ supercon- direction of the magnetic field. For such realistic geometries, ductors. Therefore, at H ϭH the homogeneous Meissner a c the concept of a demagnetization factor does not work. state is unstable in type-II superconductors and tends to split Moreover, a new type of magnetic irreversibility occurs, into normal and superconducting domains in the finest pos- which is not related to flux-line pinning but to the nonellip- sible way; this means a FLL appears with normal cores of soidal cross section that causes a ‘‘geometric barrier.’’ This radius Ϸ␰. With allowance for demagnetization effects, the barrier delays the penetration of flux lines at the four edges field of first penetration of flux lines into type-II supercon- of the rectangular cross section of the specimen. It will be ductors is thus HЈ ϭ(1ϪN)H Ͻ(1ϪN)H , and into c1 c1 c shown that this problem can be treated within a continuum type-I superconductors21 H ϭ͓(1ϪN)2H2ϩK2͔1/2Ͼ(1 p c approach, which considers the induction and current density ϪN)H , with K proportional to the wall energy. Supercon- c averaged over a few cells of the FLL. ductivity disappears when the applied field Ha reaches the ͑ ͒ ͑ ͒ critical field Hc2 type-II or Hc type-I , irrespective of de- magnetization effects, since the magnetization vanishes at 2. MAGNETIC IRREVERSIBILITY this transition. The irreversible magnetic behavior of type-II supercon- 2 The order parameter ͉␺(r)͉ and microscopic field B(r) ductors usually is caused by pinning of the Abrikosov vorti- 2 of an isolated flux line oriented along z for 2␬ ӷ1 are given ces at inhomogeneities in the material.30 However, similar 22,23 approximately by hysteresis effects have also been observed31 in type-I super- conductors, which do not contain flux lines, and in type-II ͉␺͑ ͉͒2Ϸ ͑ ϩ ␰2 2͒ r 1/ 1 2 /r , superconductors with negligible pinning. In these two cases the magnetic irreversibility is caused by a geometric ␾ ͱ 2ϩ ␰2 0 r 2 ͑specimen-shape dependent͒ barrier which delays the pen- B͑r͒Ϸ K ͩ ͪ , 2␲␭2 0 ␭ etration of magnetic flux but not its exit. In this respect the Low Temp. Phys. 27 (99–10), September–October 2001 E. H. Brandt 725

FIG. 1. Field lines of the induction B(x,y) in strips with aspect ratio b/a ϭ ͑ ͒ ϭ ͑ ͒ 2 top and b/a 0.3 bottom in a perpendicular magnetic field Ha .Top ϭ left: Ha /Hc1 0.66, in increasing field shortly before the entry field ϭ ϭ Hen /Hc1 0.665. Top right: Ha /Hc1 0.5, decreasing field. Bottom: ϭ ϭ Ha /Hc1 0.34 in increasing field just above Hen /Hc1 0.32. Note the nearly straight field lines in the corners, indicating the tension of the flux lines. The field lines of cylinders look very similar.

͑ ͒ FIG. 2. 3D plots of the screening current density Js(x,y), Eq. 11 ,in macroscopic geometric barrier behaves in a manner similar ϭ ͑ ͒ ϭ ͑ ͒ 32 superconductor strips with b/a 2 top and b/a 0.3 bottom as in Fig. 1. to the microscopic Bean–Livingston barrier for straight Ӷ Shown is the limit of small applied field Ha Hc1 before magnetic flux has vortices penetrating at a parallel surface. In both cases the penetrated. For better presentation the depicted Js(x,y) is smeared over a magnetic irreversibility is caused by the asymmetry between few grid cells. flux penetration and exit. The geometric irreversibility is most pronounced for thin films of constant thickness in a perpendicular field. It is absent only when the supercon- With increasing applied field Ha , the magnetic flux does not ϭ ductor is of exactly ellipsoidal shape or is tapered like a penetrate until an entry field Hen is reached; at Ha Hen the wedge with a sharp edge where flux can penetrate easily due flux immediately jumps to the center, from where it gradu- to the large local enhancement of the external magnetic field ally fills the entire strip or disk. This behavior in increasing at this edge in a diamagnetic material. Ha is similar to that of thin films with artificially enhanced 36,38 Ellipsoids are a particular case. In superconducting ellip- pinning near the edges, but in decreasing Ha the behav- soids the inward directed driving force exerted on the vortex ior is different: In films with enhanced edge pinning ͑critical edge͒ ends by the surface screening currents is exactly compen- current density Jc the current density J at the edge imme- 27,33 ϩ edge Ϫ edge sated by the vortex line tension. An isolated vortex line is diately jumps from Jc to Jc when the ramp rate thus in an indifferent equilibrium at any distance from the reverses its sign, while in pin-free films with a geometric specimen center. The repulsive vortex interaction therefore barrier the current density at the edge first stays constant or yields a uniform flux density, and the magnetization is re- even increases and then gradually decreases and reaches zero ͑ ϭ versible. However, in specimens of constant thickness i.e., at Ha 0. For pin-free thin strips the entry field Hen was of rectangular cross section͒ this line tension opposes the estimated in Refs. 35, 39, and 40. penetration of flux lines at the four corner lines, thus causing The outline of the present work is as follows. Section 3 an edge barrier; but as soon as two penetrating vortex seg- discusses the reversible magnetic behavior of pin-free super- ments join at the equator, they contract and are driven to the conductor ellipsoids. The effective demagnetization factor of specimen center by the surface currents; see Figs. 1 and 2. As long strips ͑or slabs͒ and circular disks ͑or cylinders͒ with opposed to this, when the specimen profile is tapered and has rectangular cross section 2aϫ2b is given in Sec. 4. In Sec. 5 a sharp edge, the driving force of the screening currents even appropriate continuum equations and algorithms are pre- in very weak applied fields exceeds the restoring force of the sented that allow one to compute the magnetic irreversibility line tension, so that there is no edge barrier. The resulting caused by pinning and/or by the geometric barrier in type-II absence of hysteresis in wedge-shaped samples was clearly superconductors of arbitrary shape, in particular, strips and shown by Morozov et al.34 disks of finite thickness. Results for thick long strips and For thin superconductor strips with an edge barrier an disks or cylinders with arbitrary aspect ratio b/a are given in elegant analytical theory of the field and current profiles has Sec. 6 for pin-free superconductors and in Sec. 7 for super- been presented by Zeldov et al.35 using the theory of com- conductors with arbitrary bulk pinning. In particular, explicit plex functions; see also the calculations in Refs. 36 and 37. expressions are given for the field of first flux entry Hen and 726 Low Temp. Phys. 27 (99–10), September–October 2001 E. H. Brandt

Ͼ for the reversibility field Hrev above which the magnetization monotonically in size for Ha Hc1 . Below for simplicity I Ӷ ͑ ␬ӷ ͒ curve is reversible and coincides with that of an ellipsoid. shall assume Hc1 Hc2 i.e., large GL parameter 1 and Ӷ Ha Hc2 . In this case one may use the model M(Ha ;0) ϭϪ ͉ ͉р Ha for Ha Hc1 and 3. ELLIPSOIDS ͑ ͒ϭ͑ ͉ ͉͉͒͑ ͉3Ϫ 3 ͒1/3Ϫ ͑ ͒ M Ha ;0 Ha / Ha Ha Hc1 Ha 6 First consider the known magnetization of ideal ellip- for ͉H ͉ϾH , which well approximates the pin-free GL soids. If the superconductor is homogeneous and isotropic, a c1 magnetization.29 the magnetization curves of ellipsoids M(Ha ;N) are revers- ible and may be characterized by a demagnetizing factor N. 4. THICK STRIPS AND DISKS IN THE MEISSNER STATE If Ha is along one of the three principal axes of the ellipsoid then N is a scalar with 0ϽNр1. One has Nϭ0 for long In nonellipsoidal superconductors the induction B(r)in specimens in a parallel field, Nϭ1 for thin films in a perpen- general is not uniform and so the concept of a demagnetizing dicular field, Nϭ1/2 for transverse circular cylinders, and factor does not work. However, when the magnetic moment ϭ ϭ ͐ ϫ 3 N 1/3 for spheres. For general ellipsoids with semi-axes a, m 1/2 r J(r)d r is directed along Ha , one may define an b, c along the Cartesian axes x, y, z, the three demagnetizing effective demagnetizing factor N which in the Meissner state ϩ ϩ ϭ ϭ ϭϪ Ϫ ͑ ͒ factors along the principal axes satisfy Nx Ny Nz 1. For (B 0) yields the same slope M/Ha 1/(1 N), Eq. 2 , ϭ ϭ ϭ ellipsoids of revolution with a b one has Nx Ny (1 as an ellipsoid with this N. Here the definition Mϭm/V with Ϫ ϭ Ͻ ϭ Nz)/2, where for ‘‘cigars’’ with a b c and for disks with m m Ha /Ha and specimen volume V is used. In particular, 18 • aϭbϾc with eccentricity eϭ͉1Ϫc2/a2͉1/2 one obtains for long strips or slabs and circular disks or cylinders with ϫ 1Ϫe2 rectangular cross section 2a 2b in a perpendicular or axial ϭ ͑ Ϫ ͒ ͑ ͒ Nz 3 arctanh e e , cigar , magnetic field along the thickness 2b, approximate expres- e ϭ sions for the slopes M/Ha m/(VHa) are given in Refs. 42 2 ϭ ͉ ͉Ϫ 1Ϫe and 43. Using this and defining q ( M/Ha 1)(b/a), one N ϭ ͑eϪarctan e͒, ͑disk͒. ͑1͒ z e3 obtains the effective N for any aspect ratio b/a in the form ϭ Ϫ ͑ ϩ ͒ For thin ellipsoidal disks with aϾbӷc one has41 N 1 1/ 1 qa/b , ␲ b a c ϭ ϩ ͫ ͩ ϩ ͪͬ N ϭ1Ϫ E͑k͒, ͑2͒ qstrip 0.64 tanh 0.64 ln 1.7 1.2 , z b 4 a b where E(k) is the complete elliptic integral of the second 4 2 b a ϭ ϩ ͫ ͩ ϩ ͪͬ ͑ ͒ 2 2 2 q tanh 1.27 ln 1 . 7 kind with k ϭ1Ϫb /a . disk 3␲ 3␲ a b When the magnetization curve in parallel field is known, ϭ ␮ Ϫ In the limits bӶa and bӷa, these formulas are exact, and M(Ha ;0) B/ 0 Ha , where B is the flux density inside Ͻ ϭ ͑ the ellipsoid, then the homogeneous magnetization of the for general b/a the relative error is 1%. For a b square cross section͒ they yield for the strip Nϭ0.538 ͑while N general ellipsoid, M(Ha ;N), follows from the implicit equa- ϭ ͒ tion 1/2 for a circular cylinder in a perpendicular field and for the short cylinder Nϭ0.365 ͑while Nϭ1/3 for a sphere͒. ϭ Ϫ ͒ ͑ ͒ Hi Ha NM͑Hi ;0 . 3 ͑ ͒ 5. COMPUTATIONAL METHOD Solving Eq. 3 for the effective internal field Hi , one ob- ϭ ϭ tains M M(Ha ;N) M(Hi ;0). In particular, for the To obtain the full, irreversible magnetization curves Meissner state (Bϭ0) one finds M(H ;0)ϭϪH and a a M(Ha) of nonellipsoidal superconductors one has to resort to numerics. Appropriate continuum equations and algo- Ha M͑H ;N͒ϭϪ for ͉H ͉р͑1ϪN͒H . ͑4͒ 44 a 1ϪN a c1 rithms have been proposed recently by Labusch and Doyle and by the author,45 based on the Maxwell equations and on ϭ ϭ Ј At the lower critical field Hc1 one has Hi Hc1 , Ha Hc1 constitutive laws which describe flux flow and pinning or ϭ Ϫ ϭ ϭϪ (1 N)Hc1 , B 0, and M Hc1 . Near the upper critical thermal depinning, and the equilibrium magnetization in ab- field Hc2 one has an approximately linear M(Ha ;0) sence of pinning, M(Ha ;0). For arbitrary specimen shape ϭ␥ Ϫ Ͻ ␥Ͼ (Ha Hc2) 0 with 0, yielding these two methods proceed as follows. ␥ While the method of Ref. 44 considers a magnetic ͒ϭ Ϫ ͒ Ϸ ͑ ͒ M͑Ha ;N ͑Ha Hc2 for Ha Hc2 . 5 charge density on the specimen surface which causes an ef- 1ϩ␥N 45 fective field Hi(r) inside the superconductor, our method ␥Ӷ ͑ ͉ ͉ Thus, if the slope 1 is small and in general, if M/Ha couples the arbitrarily shaped superconductor to the external Ӷ1 is small͒, demagnetization effects may be disregarded, field B(r,t) via surface screening currents: In a first step the Ϸ and one has M(Ha ;N) M(Ha ;0). vector potential A(r,t) is calculated for given current density The ideal magnetization curve of type II super conduc- J; then this linear relation ͑a matrix͒ is inverted to obtain J ϭ ␮ ϭ tors with N 0, M(Ha ;0) or B(Ha ;0)/ 0 Ha for given A and given Ha ; next the induction law is used to ϩ ͓ M(Ha ;0) may be calculated from Ginzburg–Landau obtain the electric field in our symmetric geometry one has t͔, and finally the constitutive law Eץ/Aץtheory,29 but to illustrate the geometric barrier any other E(J,B)ϭϪ ϭϪ ϭ model curve may be used provided M(Ha ;0) M E(J,B) is used to eliminate A and E and obtain one single Ϫ ϭ ( Ha ;0) has a vertical slope at Ha Hc1 and decreases integral equation for J(r,t) as a function of Ha(t), without Low Temp. Phys. 27 (99–10), September–October 2001 E. H. Brandt 727 having to compute B(r,t) outside the specimen. This method If one is interested also in low inductions one has to in general is fast and elegant; but so far the algorithm is generalize Eq. ͑8͒ to general reversible magnetization H restricted to aspect ratios 0.03Ͻb/aϽ30, and to a number of ϭH(B). This is achieved by replacing in the constitutive law ϭ␮Ϫ1ٌϫ grid points not exceeding 1400 ͑on a personal computer͒. E(J,B) the actual current density J 0 B by the effec- ϭٌϫ Improved accuracy is expected by combining the methods of tive current density JH H which drives the vortices and ϭٌϫ Refs. 44 ͑working best for small b/a͒ and 45. Here I shall thereby generates an electric field E. That JH H(B,r) use the method of Ref. 45 and simplify it to the two- enters the Lorentz force is rigorously proven by Labusch.44 dimensional ͑2D͒ geometry of thick strips and disks. Within the London theory this important relation may also be In the 2D geometry of thick strips42 or short cylinders43 inferred from the facts that the force on a vortex is deter- ϭ␮ ϭٌϫ in an applied magnetic field Ba 0Ha Aa along y, one mined by the local current density at the vortex center, while writes rϭ(x,y)orrϭ(␳,y) ͑in cylindrical coordinates ␳, ␸, the energy density F of the vortex lattice is determined by ϭٌ y͒. For a uniform applied field the applied vector potential in the magnetic field at the vortex centers. Thus, JH ץ ץ ϭϪ ϭϪ␳ ϫ these two geometries reads Aa xBa or Aa Ba/2. The ( F/ B) is the average of the current densities at the vor- current density J(r,t), electric field E(r,t), and vector po- tex centers, which in general is different from the current ϭ␮Ϫ1ٌϫ tential A(r,t) now have only one component oriented along density J 0 B averaged over the vortex cells. In our z or ␸ and denoted by J, E, A. The method42,43,45 describes 2D geometry one thus has to replace in Eq. ͑8͒ the superconductor by its current density J(r,t), from which ͓ ͑ Ј͒ ͑ Ј͔͒→ ͓ ͑ Ј͒ ͑ Ј͔͒ ͑ ͒ ϭ ␳ E J r ,B r E JH r ,B r , 12 the magnetic field B(x,y,t) (Bx ,By)orB( ,y,t) ϭ ͑ ͒ y depends on the reversibleץ/ HץxϪץ/ HץB␳ ,By), the magnetic moment m(t) along y , and the where J ϭ) ϭ Ј H y x B with H ϭH(B)B /B, Hץ/Fץelectric field E(r,t) E(J,B,r ) follow directly or via the constitutive law H(B)ϭ ϭ x x y constitutive law E E(J,B). For high inductions B ϭH(B)B /B, and Bϭ(B2ϩB2)1/2. ӷ␮ Ϸ␮ ϭϪ␮Ϫ1ٌ2 y x y 0Hc1 one has B 0H everywhere and J 0 (A The boundary condition on H(r) is simply that one has Ϫ Aa). The current density J is then obtained by time- ϭ ␮ ͑ H B/ 0 at the surface and in the vacuum outside the su- integrating the following equation of motion: perconductor, which does not enter our calculation͒. This 1 boundary condition may be forced by an appropriate space- ͑ ͒ϭϪ ͵ 2 Ј ͑ Ј͓͒ ͑ ͒ϩ ˙ ͑ Ј ͔͒ ͑ ͒ ϭ J r,t ␮ d r K r,r E J,B Aa r ,t . 8 dependent constitutive law H H(B,r), which outside and at 0 V ϭ ␮ the surface of the superconductor is trivially H B/ 0 . The Ј ϭ Ј Ϫ1 specimen shape thus enters in two places: via the integral Here K(r,r ) Q(r,r ) is an inverse integral kernel ob- Ј ϭ tained by inverting a matrix; see Refs. 42 and 43 for details. kernel K(r,r ) and via the constitutive law H H(b,r). The kernels Q and K apply to the appropriate geometry and To compute the induction B(r) entering H(B), for maxi- Ϫ mum accuracy one should not use the derivative BϭٌϫA relate J to the current-caused vector potential A Aa in the ϭ but the Biot–Savart integral ٌ ͒ ͑ here trivial gauge •A 0 via integrals over the specimen volume V, ͑ ͒ϭ ͵ 2 Ј ͑ Ј͒ ͑ Ј͒ϩ ͑ ͒ ͑ ͒ B r d r L r,r J r Ba r 13 V ͑ ͒ϭ␮ ͵ 2 Ј ͑ Ј͒ ͑ Ј͒ϩ ͑ ͒ ͑ ͒ A r 0 d r Q r,r J r Aa r , 9 V with a suitable kernel L(r,rЈ). The accuracy of the method then depends mainly on the algorithm used to compute the 1 ϭ ͵ 2 Ј ͑ Ј͓͒ ͑ Ј͒Ϫ ͑ Ј͔͒ ͑ ͒ derivative J ϭٌϫH. A useful trick is to compute J as͒ ͑ J r ␮ d r K r,r A r Aa r . 10 H H V ϭ ϩٌϫ Ϫ ␮ Ϫ ␮ 0 JH J (H B/ 0), where H B/ 0 is typically small and vanishes at the surface. The Laplacian kernel Q is universal, e.g., Q(r,rЈ) For the following computations I use simple models for ϭϪ(1/2␲)ln͉rϪrЈ͉ for long strips with arbitrary cross sec- the constitutive laws of an isotropic homogeneous type-II tion, but the inverse kernel K depends on the shape of the superconductor without the Hall effect, though our method45 specimen cross section. Putting A(rЈ)ϭ0 in Eq. ͑10͒͑Meiss- is more general. With Eq. ͑6͒ and HϭB/␮ ϪM one has ner state͒, one sees that 0 Ϫ H͑B͒ϭ␮ 1͓B3 ϩB3͔1/3 ͑14͒ 1 0 c1 J ͑r͒ϭϪ ͵ d2rЈK͑r,rЈ͒A ͑rЈ͒ ͑11͒ s ␮ a ϭ␮ 0 V with Bc1 0Hc1 . A simple B-dependent current–voltage law which describes pinning, thermal depinning, and flux is the surface screening current caused by the applied field. flow is E(J,B)ϭ␳(J,B)J, with ϭ In particular, one has Js(r) 0 inside the superconductor. In our above method J automatically is restricted to the layer ͑J/J ͒␴ s ␳͑ ͒ϭ␳ c ͑ ͒ of grid points nearest to the surface; see Fig. 2. Analytical J,B 0B ϩ ͒␴ . 15 1 ͑J/Jc expressions for the current Js in thick rectangular strips with ␳ϰ ␴ ͑ Ӷ ͒ applied field Ha and/or applied current Ia were recently This model has the correct limits J , J Jc , flux creep 46 ␳ϭ␳ ϭ␳ ͑ ӷ ␳ ϭ ͒ given for this limit of vanishing magnetic penetration depth and 0B FF J Jc , flux flow, 0 const , and for ␭→0. Finite ␭Ͼ0 may be introduced into these computa- large creep exponent ␴ӷ1 it reduces to the Bean critical tions by modifying the integral kernel according to Ref. 47: state model. In general the critical current density Jc ϭ ϩ␭2␦ Ϫ Ϫ1 ϭ ␴ у K(r,rЈ) ͓Q(r,rЈ) (r rЈ)͔ . The resulting screen- Jc(B) and the creep exponent (B) 0 will depend on B. ␭ → ing current then flows in a surface layer of finite thickness . For pin-free superconductors (Jc 0) this expression de- 728 Low Temp. Phys. 27 (99–10), September–October 2001 E. H. Brandt

Ϫ FIG. 4. The irreversible magnetization curves M(Ha) of pin-free circular disks and cylinders with aspect ratios b/aϭ0.08, 0.15, 0.25, 0.5, 1, 2, 5, and ϱ in an axial field ͑solid lines͒. Here the irreversibility is due only to a geometric edge barrier for flux penetration. The reversible magnetization curves of the corresponding ellipsoids defined by Eqs. ͑3͒, ͑6͒,and͑7͒ are shown as dashed lines.

lar to pinning, causes a weak hysteresis and a small remanent ϭ flux at Ha 0; this artifact is reduced by choosing a larger ϭ FIG. 3. The axial magnetic induction By(r,y) in the midplane y 0ofa resistivity or a slower ramp rate. pin-free superconductor disk with aspect ratio b/aϭ0.3 in increasing field In Fig. 3 the induction B ϭB (0,0) in the specimen cen- ͑ ͒ ͑ ͒ 0 y solid lines and then decreasing field dashed lines , plotted at Ha /Hc1 ter performs a hysteresis loop very similar to the magnetiza- ϭ0.4, 0.42, ..., 0.5, 0.52, 0.6, 0.7, 0.8, 0.7, 0.6, ..., 0.1, 0 ͑a͒. The induction B (0,0) as the center of the same disk ͑solid line͒ and of a strip ͑dashed tion loops M(Ha) shown in Figs. 3 and 4. Both loops are y Ϫ ϭϪ Ϫ line͒, both with b/aϭ0.3. The symbols mark the field values at which the symmetric, M( Ha) M(Ha) and B0( Ha) profiles are taken. Also shown are the magnetization loops for the same disk ϭϪ B0(Ha). The maximum of M(Ha) defines a field of first and strip and the corresponding reversible magnetization ͑dotted lines͒͑b͒. Ј flux entry Hen , which closely coincides with the field Hen at which By(0,0) starts to appear. The computed entry fields are well fitted by scribes usual flux flow, i.e., viscous motion of vortices, E strip ϭ ͱ ϭ␳ ␳ ϰ Hen /Hc1 tanh 0.36b/a, FF(B)J, with the flux-flow resistivity FF B,asit should be. disk ϭ ͱ ͑ ͒ Hen /Hc1 tanh 0.67b/a. 16 These formulas are good approximations for all aspect ratios Ͻ Ͻϱ Ϸͱ 6. PIN-FREE SUPERCONDUCTORS 0 b/a ; see also the estimates of Hen b/a for thin strips in Refs. 35 and 39. The penetration and exit of flux computed from Eqs. The virgin curve of the irreversible M(Ha) of strips and ͑ ͒ ͑ ͒ 8 – 15 is visualized in Figs. 1–3 for isotropic strips and disks at small Ha coincides with the ideal-Meissner straight ϭϪ Ϫ disks without volume pinning, using a flux-flow resistivity line M Ha /(1 N) of the corresponding ellipsoid, Eqs. ␳ ϭ␳ ␳ ϭ ͑ ͒ ␳ ϭ ͑ ͒ ͑ ͒ ͑ ͒ FF 0B(r) with 0 140 strip or 0 70 disk , in units 4 , 7 . When the increasing Ha approaches Hen , flux starts ϭ ϭ␮ ϭ͉ ͉ϭ where Hc1 a 0 dHa /dt 1. Figure 1 shows the field to penetrate into the corners in the form of almost straight ͑ ͒ ͉ ͉ lines of B(x,y) in two pin-free strips with aspect ratios b/a flux lines Fig. 1 , and thus M(Ha) falls below the Meiss- ϭ ϭ ϭ 2 and b/a 0.3; Fig. 2 shows the surface screening cur- ner line. At Ha Hen flux penetrates and jumps to the center, ͉ ͉ rents in the same strips before flux has penetrated; and Fig. 3 and M(Ha) starts to decrease. In decreasing Ha , this bar- plots some induction profiles in a strip and some hysteresis rier is absent. As soon as flux exit starts, all our calculated loops of the magnetization and of the induction at the center M(Ha) exhibit strong ‘‘numerical noise,’’ which reflects the of a strip and disk. instability of this state. Similar but weaker noise occurs at The profiles of the induction By(r,y) taken along the the onset of flux penetration. ϭ midplane y 0 of the thick disk in Fig. 3 have a pronounced As can be seen in Fig. 4, above some field Hrev , the ϭ minimum near the edge r a, which is the region where magnetization curve M(Ha) becomes reversible and exactly strong screening currents flow. Away from the edges, the coincides with the curve of the ellipsoid defined by Eqs. ͑3͒, ͒ → ϭٌϫ ␮ ͑ ͒ ͑ ͒͑ ␳Ϫ1 current density J B/ 0 is nearly zero; note the parallel 6 , and 7 in the quasistatic limit with 0 dHa /dt 0 . ϭٌϫ field lines in Fig. 1. The quantity JH H(B), which en- The irreversibility field Hrev is difficult to compute since it ϫ ␳ ters the Lorentz force density JH B, is even exactly zero, depends slightly on the choices of the flux-flow parameter 0 since we assume the absence of pinning and the viscous drag ͑or ramp rate͒ and of the numerical grid, and also on the ␳ р Ͻ force is small. Our finite flux-flow parameter 0 and finite model for M(Ha ;0). In the interval 0.08 b/a 5 we find, ϭϮ ramp rate dHa /dt 1 mean a dragging force which, simi- with relative error of 3%, Low Temp. Phys. 27 (99–10), September–October 2001 E. H. Brandt 729

Ϫ ϭϪ FIG. 5. The magnetization curves M( Ha) M(Ha) of a thick disk with FIG. 7. Magnetization curves of the same disk as in Fig. 5 but for the Kim ϭ aspect ratio b/a 0.25 and constant Hc1 for various pinning strengths, Jc model, J (B)ϭJ /(1ϩ3͉B͉/B ) for various pinning strengths J ϭ0, ϭ c c0 c1 c0 0, 0.25, 0.5, 1, 1.5, 2, 3, 4 in units of Hc1 /a, and various sweep ampli- 0.25, 0.5, 1, 1.5, 2, 3, 4 in units of H /a. Presentation as in Fig. 5. ϭ c1 tudes. Bean model. The inner loop belongs to the pin-free disk (Jc 0), the outer loop to strongest pinning. The reversible magnetization curve of the corresponding ellipsoid is shown as a dashed curve. below Hc1 . The maximum magnetization may be estimated from the slope of the virgin curve 1/(1ϪN), Eq. ͑7͒, and strip ϭ ϩ ͑ ͒ from the field of first flux entry, Eq. ͑16͒. Hrev /Hc1 0.65 0.12 ln b/a , Formulas ͑7͒, ͑16͒, and ͑17͒ are derived essentially from disk ϭ ϩ ͑ ͒ ͑ ͒ Hrev /Hc1 0.75 0.15 ln b/a . 17 first principles, with no assumptions but the geometry and This fit obviously does not apply to very small b/aӶ1 ͑since finite Hc1 . They should be used to interpret experiments on Ͼ ͒ ӷ ͑ superconductors with no or very weak vortex pinning. At Hrev should exceed Hrev 0 nor to very large b/a 1 where ͒ present it is not clear how the presence of a microscopic Hrev should be close to Hc1 . The limiting value of Hrev for thin films with bӶa is thus not yet known. Bean–Livingston barrier may modify these continuum theo- Remarkably, the irreversible magnetization curves retical results. M(Ha) of pin-free strips and disks fall on top of each other if the strip is chosen twice as thick as the disk, (b/a)strip 7. SUPERCONDUCTORS WITH PINNING Ϸ2(b/a) . This striking coincidence holds for all aspect disk Figures 5–8 show how the irreversible magnetization ratios 0Ͻb/aϽϱ and can be seen from each of Eqs. ͑7͒, loops of disks with b/aϭ0.25 ͑and in Fig. 9 for a thinner ͑16͒, and ͑17͒. The effective N ͓or virgin slope 1/(1ϪN)͔, disk with b/aϭ0.125͒ are modified when volume pinning is the entry field H , and the reversibility field H are nearly en rev switched on. In Figs. 5, 6, and 9, pinning is described by the equal for strips and disks with half thickness, or for slabs and Bean model with constant critical current density J , while cylinders with half length. c in Figs. 7 and 8 the Kim model is used with an induction- Another interesting feature of the pin-free magnetization ϭ ϩ ͉ ͉ ϭ␮ ͉ ͉ dependent Jc(B) Jc0 /(1 3 B /BK), with BK 0Hc1/3 loops is that the maximum of M(Ha) exceeds the maxi- ͑ ͒ Ͻ mum of the reversible curve equal to Hc1 when b/a 0.8 for strips and b/aϽ0.4 for disks, but at larger b/a it falls

FIG. 8. Magnetization curves as in Fig. 6 but for the Kim model Jc(B) ϭ ϩ ͉ ͉ ϭ ϭ Jc0 /(1 3 B /aJc0) with Jc0 const for various Hc1 0, 0.1, 0.2, 0.35, 0.5, 0.7, 1 in units of aJc0 . Also depicted are the pin-free magnetization ϭ ͑ ϭ ͒ FIG. 6. Magnetization curves of a disk as in Fig. 5 but with Jc const and line with dots; M and Ha here are in units Hc1 since Jc0 0 and the for various lower critical fields Hc1 in units of aJc . Bean model. irreversible magnetization of the corresponding ellipsoid. 730 Low Temp. Phys. 27 (99–10), September–October 2001 E. H. Brandt

plane anisotropy of pinning occurs, e.g., in high-Tc super- conductors with layered structure.

*E-mail: [email protected]

1 H. Kamerlingh Onnes, Leiden Comm. 120b, 122b, 124c, ͑1911͒. 2 J. K. Hulm and B. T. Matthias, Science 208,881͑1980͒. 3 J. G. Bednorz and K. A. Mu¨ller, Z. Phys. B64,189͑1986͒. 4 M. K. Wu, J. R. Ashburn, C. J. Torng, P. H. Hor, R. L. Meng, L. Gao, Z. J. Huang, Y. Q. Wang, and C. W. Chu, Phys. Rev. Lett. 58,908͑1987͒. 5 H. Maeda, Y. Tanaka, M. Fukutomi, and T. Asano, Jpn. J. Appl. Phys. 27, L206 ͑1988͒. 6 S. X. Dou, X. L. Wang, Y. C. Guo, Q. Y. Hu, P. Mikheenko, J. Horvath, M. Ionescu, and H. K. Liu, Supercond. Sci. Technol. 10, A52–A67 ͑1997͒. FIG. 9. Same magnetization curves as in Fig. 5 but for a thinner disk with 7 Z. Z. Sheng and A. M. Hermann, Nature ͑London͒ 332,55͑1988͒. ϭ ϭ aspect ratio b/a 0.125 for various degrees of pinning Jca/Hc1 0, 0.25, 8 L. Gao, Y. Y. Xue, F. Chen, Q. Xiong, R. L. Meng, D. Ramirez, C. W. 0.5, 1, 1.5, 2, 3, 4 and constant Hc1 . Chu, J. H. Eggert, and H. K. Mao, Phys. Rev. B 50, 4260 ͑1994͒. 9 Y. C. Kim, J. R. Thompson, J. G. Ossandon, D. K. Christen, and M. Paranthaman, Phys. Rev. B 51, 11767 ͑1995͒. ͑ ͒ ϭ␮ ͑ ͒ 10 J. Nagamatsu, N. Nakagawa, T. Muranaka, Y. Zenitani, and J. Akimitsu, Fig. 8 or BK 0aJc0/3 Fig. 9 . In Figs. 5, 7, and 9, Hc1 is ͑ ͒ ͑ ͒ ͑ Nature London 410,63 2001 . held constant; with increasing Jc or Jc0 in natural units 11 Magnetic Properties and Critical Currents of Superconducting Alloys,J. ͒ Hc1 /a the magnetization loops are inflated nearly sym- N. Rjabinin and L. W. Schubnikow, Phys. Z. Sowjetunion 7, N1, 122 metrically about the pin-free loop or about the reversible ͑1935͒; Nature ͑London͒ 135, 581 ͑1935͒. 12 curve ͑above H ͒, and the maximum of ͉M(H )͉ shifts to L. W. Schubnikow, W. I. Chotewitsch, J. D. Schepelew, and J. N. Rjabinin, rev a Phys. Z. Sowjetunion 10, N2, 39 ͑1936͒; Zh. E´ ksp. Teor. Fiz. 7,221 higher fields. Above Hrev the width of the loop is nearly ͑1937͒. 42,43 13 ´ ͑ ͓͒ proportional to Jc , as expected from theories which as- A. A. Abrikosov, Zh. Eksp. Teor. Fiz. 32, 1442 1957 Sov. Phys. JETP 5, ϭ 1174 ͑1957͔͒. sume Hc1 0, but at small fields the influence of finite Hc1 is 14 V. L. Ginzburg and L. D. Landau, Zh. E´ ksp. Teor. Fiz. 20, 1064 ͑1950͒ clearly seen up to rather strong pinning. ͓English translation in Men of Physics͔, edited by L. D. Landau and D. ter In Figs. 6 and 8, Jc or Jc0 is held constant and Hc1 Haar, New York: Pergamon ͑1965͒, Vol. 1, p. 138. ͑ ͒ 15 ͑ ͒ increased from zero in the natural units aJc . As expected, J. Pearl, Appl. Phys. Lett. 5,65 1964 . 16 A. L. Fetter and P. C. Hohenberg, Phys. Rev. 159, 330 ͑1967͒. the influence of finite Hc1 is most pronounced at small ap- 17 Ϫ J. R. Clem, Phys. Rev. B 43, 7837 ͑1991͒. plied fields Ha , where it causes a peak in M even in the 18 L. D. Landau and E. M. Lifshitz, Electrodynamics of Continuous Media, Bean magnetization curves, which without consideration of Vol. 8, in Theoretical Physics, Pergamon Press, London ͑1959͒. 19 ͑ ͒ Hc1 consist of two monotonic branches and a monotonic A. Hubert, Phys. Status Solidi 24,669 1967 . 20 ͑ ͒ virgin curve. Within the Kim model, or with any decreasing A. Fortini and E. Paumier, Phys. Rev. B 5, 1850 1972 . 21 H. Kronmu¨ller and H. Riedel, Phys. Status Solidi 77, 581 ͑1976͒. Jc(B) dependence, the magnetization loops exhibit a maxi- 22 J. R. Clem, J. Low Temp. Phys. 18,427͑1975͒. ϭ 48 23 mum even when Hc1 0 is assumed. With increasing Hc1 Z. Hao, J. R. Clem, M. W. McElfresh, L. Civale, A. P. Malozemoff, and F. this maximum becomes sharper and shifts to larger fields ͑cf. Holtzberg, Phys. Rev. B 43, 2844 ͑1991͒. ͒ 24 M. Tinkham, Introduction to Superconductivity, McGraw-Hill, New York Fig. 8 . Comparing Figs. 5 and 9, one sees that for supercon- ͑ ͒ Ͼ 1975 . ductor disks with pinning and with Hc1 0, the peak in 25 E. H. Brandt, J. Low Temp. Phys. 26, 709 ͑1977͒; ibid. ,735͑1977͒; ibid. Ϫ ͑ ͒ ͑ ͒ M(Ha) becomes more pronounced and shifts towards 28,263 1977 ; Phys. Rev. B 34, 6514 1986 . smaller applied fields when the disk thickness is decreased. 26 E. H. Brandt and U. Essmann, Phys. Status Solidi B 144,13͑1987͒. 27 E. H. Brandt, Rep. Prog. Phys. 58,1465͑1995͒. In the classical Bean model, i.e., if the lower critical field 28 ͑ G. Blatter, M. V. Feigel’man, V. B. Geshkenbein, A. I. Larkin, and V. M. Hc1 and the B dependence of Jc(B) are disregarded both Vinokur, Rev. Mod. Phys. 66, 1125 ͑1994͒. conditions are satisfied when B is sufficiently high͒, there 29 E. H. Brandt, Phys. Rev. Lett. 78, 2208 ͑1997͒. 30 P. W. Anderson, Phys. Rev. Lett. 9, 309 ͑1962͒. exist analytical solutions for the critical state not only for the 31 32 J. Provost, E. Paumier, and A. Fortini, J. Phys. F: Met. Phys. 4,439 simple longitudinal geometry but also for the more realistic ͑1974͒; A. Fortini, A. Haire, and E. Paumier, Phys. Rev. B 21, 5065 49 geometries of thin disks in an axial field and for long thin ͑1980͒. strips in a perpendicular field.50 Interestingly, the expressions 32 C. P. Bean and J. D. Livingston, Phys. Rev. Lett. 12,14͑1964͒; for the profiles of the current density, J(␳) and J(x), have L. Burlachkov, Phys. Rev. B 47, 8056 ͑1993͒. 33 M. V. Indenborn, H. Kronmu¨ller, T. W. Li, P. H. Kes, and A. A. Menovsky, identical form in these two geometries, but an analytical ex- Physica C 222, 203 ͑1994͒; M. V. Indenbom and E. H. Brandt, Phys. Rev. ␳ pression for the magnetic field profiles, By( ) and By(x), Lett. 73, 1731 ͑1994͒. exists only for the strip geometry but not for the disk. Re- 34 N. Morozov, E. Zeldov, D. Majer, and B. Khaykovich, Phys. Rev. Lett. 76, 138 ͑1996͒; N. Morozov, E. Zeldov, V. Konczykowski, and R. A. Doyle, cently the critical-state problem has been solved also for thin ͑ ͒ 41 Physica C 291,113 1997 . ellipsoidal disks in a perpendicular magnetic field; this gen- 35 E. Zeldov, A. I. Larkin, V. B. Geshkenbein, M. Konczykowski, D. Majer, eral solution contains the circular disk and long strip as lim- B. Khaykovich, V. M. Vinokur, and H. Strikhman, Phys. Rev. Lett. 73, iting cases. Exact solutions were also obtained when the 1428 ͑1994͒. 36 critical current density in thin films depends on the orienta- E. Zeldov, A. I. Larkin, M. Konczykowski, B. Khaykovich, D. Majer, V. B. Geshkenbein, and V. M. Vinokur, Physica C 235–240,2761͑1994͒; tion of the local magnetic field with respect to the film plane, B. Khaykovich, E. Zeldov, M. Konczykowski, D. Majer, A. I. Larkin, and i.e., on the inclination angle of the flux lines.51 This out-of- John R. Clem, ibid. 235–240, 2757 ͑1994͒. Low Temp. Phys. 27 (99–10), September–October 2001 E. H. Brandt 731

37 I. L. Maksimov and A. A. Elistratov, JETP Lett. 61, 208 ͑1995͒;I.L. 45 E. H. Brandt, Phys. Rev. B 59, 3369 ͑1999͒. Maximov and A. A. Elistratov, Appl. Phys. Lett. 72, 1650 ͑1998͒. 46 E. H. Brandt and G. P. Mikitik, Phys. Rev. Lett. 85, 4164 ͑2000͒. 38 Th. Schuster, M. V. Indenborn, H. Kuhn, E. H. Brandt, and M. 47 E. H. Brandt, Phys. Rev. B 64, 24505 ͑2001͒. Konczykowski, Phys. Rev. Lett. 73,1424͑1994͒. 48 D. V. Shantsev, M. R. Koblischka, Y. M. Galperin, T. H. Johansen, L. Pust, 39 ͑ ͒ M. Benkraouda and J. R. Clem, Phys. Rev. B 53, 5716 1996 ; ibid. 58, and M. Jirsa, Phys. Rev. Lett. 82,2947͑1999͒. 15103 ͑1998͒. 49 ͑ ͒ 40 P. N. Mikheenko and Yu. E. Kuzovlev, Physica C 204,229 1993 . A. V. Kuznetsov, D. V. Eremenko, and V. N. Trofimov, Phys. Rev. B 56, 50 E. H. Brandt, M. V. Indenborn, and A. Forkl, Europhys. Lett. 22,735 9064 ͑1997͒; ibid. 57, 5412 ͑1998͒. ͑1993͒. 41 G. P. Mikitik and E. H. Brandt, Phys. Rev. B 60, 592 ͑1999͒. 51 G. P. Mikitik and E. H. Brandt, Phys. Rev. B 62, 680 ͑2000͒. 42 E. H. Brandt, Phys. Rev. B 54, 4246 ͑1996͒. 43 E. H. Brandt, Phys. Rev. B 58, 6523 ͑1998͒. 44 R. Labusch and T. B. Doyle, Physica C 290, 143 ͑1997͒; T. B. Doyle, This article was pubblished in English in the original Russian journal. Re- R. Labusch, and R. A. Doyle, ibid. 290,148͑1997͒. produced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Vortex matter in superconductors V. M. Pan*

G. V. Kurdyumov Institute of Metal Physics, National Academy of Sciences of Ukraine, bul’var Vernadskogo 36, 03142 Kiev, Ukraine A. V. Pan**

G. V. Kurdyumov Institute of Metal Physics, National Academy of Sciences of Ukraine, bul’var Vernadskogo 36, 03142 Kiev, Ukraine; Institute for Superconducting and Electronic Materials, University of Wollongong, New South Wales 2522, Australia ͑Submitted May 18, 2001͒ Fiz. Nizk. Temp. 27, 991–1011 ͑September–October 2001͒ The behavior of the ensemble of vortices in the Shubnikov phase in biaxially oriented films of ͑ ͒ the high-temperature superconductor YBa2Cu3O7Ϫ␦ YBCO in an applied magnetic field is investigated for different orientations of the field. The techniques used are the recording of the current–voltage characteristics in the transport current and of resonance curves and damping of a mechanical oscillator during the passage of a transport current. It is shown that the behavior of the vortex ensemble in YBCO films, unlike the case of single crystals, is determined by the interaction of the vortices with linear defects—edge dislocations, which are formed during the pseudomorphic epitaxial growth and are the dominant type of defect of the crystal lattice, with a density reaching 1015 lines/m2. The effective pinning of the vortices ͑ у ϫ 10 2 ͒ and the high critical current density Jc 3 10 A/m at 77 K in YBCO films are due precisely to the high density of linear defects. New phase states of the vortex matter in YBCO films are found and are investigated in quasistatics and dynamics; they are due to the interaction of the vortices with crystal defects, to the onset of various types of disordering of the vortex lattice, and to the complex depinning process. A proposed H – T phase diagram of the vortex matter for YBCO films is proposed. © 2001 American Institute of Physics. ͓DOI: ͓DOI: 10.1063/1.1401182͔

INTRODUCTION state in films of the moderately anisotropic high-temperature ͑ ͒ ͑ ͒ superconductor HTSC YBa2Cu3O7Ϫ␦ YBCO . It was first shown by the Ukrainian physicist L. V. Shub- nikov back in 19361 that in superconducting alloys there ex- 1. QUANTIZED VORTICES IN A SUPERCONDUCTOR ists a wide range of magnetic fields at which the Meissner effect gradually diminishes and the magnetic flux penetrates The penetration of magnetic field into a type-II super- into the volume of the superconductor. In that case the re- conductor occurs in the form of quantized vortex lines or sponse of the superconductor to an increase in the external flux lines. Each such flux line carriers a quantum of magnetic ͑ ͒ ⌽ ϭ ϭ ϫ Ϫ15 2 magnetic field i.e., the magnetization curves have a com- flux, 0 hc/2e 2.07 10 T•m , and has a normal pletely different form than in the case of pure metals, i.e., core, which in an isotropic superconductor is a thin normal type-I superconductors. The term ‘‘type-II superconductor’’ cylinder along the magnetic field. The radius of this cylinder, was first used for alloys and compounds by Abrikosov in the vortex core, is equal to the coherence length ␰ ͑an impor- 1957 in developing a consistent theory of type-II tant scale length in a superconductor, which in the micro- superconductors.2 This theory made it possible to understand scopic theory is defined as the distance between interacting the experimental results of Shubnikov on the basis of the electrons in a Cooper pair and in the Ginzburg–Landau ͑GL͒ concepts of flux quantization and the penetration of magnetic phenomenological theory, as the distance at which the super- field into type-II superconductors in the form of a lattice of conducting order parameter varies from its maximum to zero Abrikosov vortices. On the H – T phase diagram ͑i.e., the at a superconductor/normal metal boundary͒.3 Around the diagram in magnetic field versus temperature͒ the phase in normal core flows an undamped supercurrent, which in iso- which a type-II superconductor is threaded by vortices— tropic type-II superconductors is oriented in such a way that magnetic flux quanta—has ever since that time been called the magnetic field induced by it is directed along the core the Shubnikov phase. As time went on, it became clear that and coincides with the direction of the external field. The vortex states in superconductors are very complex and di- vortex current flows in a region with a radius of the order of verse while at the same time being exceedingly important for ␭, the London penetration depth of a weak magnetic field. understanding the behavior of superconductors in an electro- For a type-II superconductor this region is much larger than ␰ ␴ Ͻ magnetic field and under current loading. A new field of , and this is a consequence of the fact that ns 0ina physics has arisen, which might be called vortex-matter type-II superconductor, i.e., ␭ӷ␰. The penetration of vorti- physics. In this article we examine the features of the vortex ces into a type-II superconductor becomes energetically fa-

1063-777X/2001/27(9–10)/15/$20.00732 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan 733

Ͼ ͑ vorable in an external field Hext Hc1 Hc1 is the first, or shown to strongly alter the form of this phase diagram, add- lower, critical field͒. Penetrating into the volume of the ing new features and new phase regions. type-II superconductor, the vortices are spaced a distance ϰ ⌽ 1/2 р␭ a0 ( 0 /H) apart, and when a0 they begin to interact ͑ ͒ ϭ repelling each other owing to the Lorentz force fL 1/c ͓j 3. THE H – T PHASE DIAGRAM IN THE PRESENCE OF ϫ⌽ 0͔, forming a regular triangular lattice in the transverse CORRELATED DISORDERING cross section.3,4 The presence of linear defects, which by their nature are capable of pinning vortices most strongly along their entire 2. VORTEX STATE—A NEW FORM OF CONDENSED length and whose density is comparable to the density of MATTER vortices will give rise to a special state, the so-called ‘‘Bose glass,’’ when the vortices are localized in a random manner A powerful impulse for the further development of the by linear defects lying parallel to one another. The mecha- physics of the vortex state was obtained after the discovery nism of formation of the two-dimensional ͑or correlated͒ of high-Tc superconductors in 1986. This was because of two Bose glass is reminiscent of the mechanism of formation of important circumstances: first, the critical temperature of the three-dimensional vortex glass ͑which is another disor- HTSC cuprates is so high that they become superconducting dered state of the vortex lattice, in which it interacts with a at temperatures where thermal fluctuations play an appre- set of random pointlike defects of the crystal͒:6–10 the linear ciable role, since their energy becomes comparable to the defects also generate additional low-lying vortex states elastic energy of a vortex and/or of the vortex lattice and also which differ from those already existing in the perfect lattice, to the pinning energy, thus creating the necessary conditions and the vortices are trapped by these states if the potential for the appearance of unusual new regions on the H – T phase wells induced by the disorder are sufficiently deep in com- diagram of the superconductor—different states of the vortex parison with the energy of thermal fluctuations. The charac- matter and phase transitions between them; second, owing to ter of the H – T phase diagram, as in the case of pointlike the layered crystal structure and the anisotropy inherent to disorder, is determined by the compensation and competition HTSC metal-oxide cuprates, favorable conditions are created of the elastic, pinning, and thermal characteristic energies for the appearance on the H – T diagram of phase regions and and also by the use of the corresponding Lindemann phase transitions involving changes in the dimensionality in criterion.5 the vortex ensemble from three-dimensional to two- The upward shift of the melting line on the H – T dia- dimensional and vice versa. gram in the presence of linear defects can be estimated, The vortex matter in type-II superconductors is a unique keeping in mind that the modulation of the parabolic elastic example of a condensed state with controllable parameters.5 potential of the vortex lattice by the linear pinning potential Unlike ordinary condensed-matter systems the density of the causes an increase in its effective depth. Therefore, in order constituent particles ͑magnetic vortices͒ and their interaction to remove a given vortex from the confines of the ‘‘box,’’ the can be changed by several orders of magnitude in a control- thermal fluctuations must overcome not only the elastic bar- lable way by simply varying the external magnetic field. In rier but also additional pinning barriers. Thus the condition addition, extremely important thermal fluctuation effects in for the loss of stability of the vortex lattice can be deter- experiments with HTSC cuprates can be observed over a mined from the equation wide temperature range, and these effects are reflected on the k TϭE ϩE , ͑1͒ H – T phase diagram. Finally, vortex matter is the most con- B elastic pin venient tool for studying disordered media—one of the cen- where kBT, Eelastic , and Epin are the characteristic energies of tral problems of condensed-matter physics. thermal fluctuations, the elastic interaction, and pinning. The In a conventional treatment of the Shubnikov vortex form of Eq. ͑1͒ shows that linear defects are the cause of the state without allowance for thermal fluctuation and pinning shift of the solid↔liquid transition, which in this case is effects, it is assumed that a homogeneous solid vortex-lattice called the Bose glass transition, HBG(T), to higher tempera- phase exists in the field interval between the lower critical tures in comparison with the melting temperature of the per- field Hc1 , where the vortices begin to penetrate into the su- fect crystal in the absence of these linear defects, HBG(T) Ͼ 11 perconductor, and the mean-field upper critical field Hc2 , Hm(T). In the case of a fairly low density of linear de- above which the superconductivity vanishes. fects, when the main contribution is given by the elastic en- The upper critical field Hc2(T) of a type-II supercon- ergy and the disordering leads only to small shifts of the ductor is defined as the field at which, at a temperature T, the melting line, the melting of the Bose glass still preserves the normal cores of the vortices begin to overlap, and the super- features of a first-order transition. Such a transition occurs in Ͻ ϭ⌽ 2 conductor becomes ‘‘normal.’’ If one ignores thermal fluc- low magnetic fields H H⌽ , where H⌽ 0 /D is the so- ϭ⌽ ␲␰2 tuations, then Hc2 0 /(2 ), and this means that a small called ‘‘matching’’ field, at which the density of linear de- ͑ coherence length is conducive to an increase in Hc2 . fects equals the density of vortices D is the average distance ␰ ͒ Ͼ For high-Tc cuprate superconductors the small value of between linear defects . At higher fields H H⌽ the correc- is the reason why Hc2 far exceeds 100 T at low temperatures. tions to the melting temperature rapidly decrease, and the The well-known H – T phase diagram for a ‘‘clean’’ or ideal melting line of the Bose glass returns to its original position. type-II superconductor can be obtained using the ‘‘mean field Thus the greatest deviation of the melting line of the Bose approximation’’ in the GL theory. However, taking thermal glass from its original position in perfect crystals is expected fluctuations and disordering effects into account has been to occur near HХH⌽ . 734 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan

FIG. 1. Schematic H – T phase diagram of the vortex matter in the presence of statistically distributed extended linear ͑columnar͒ defects.5

Figure 1 shows a schematic phase diagram of a vortex system with correlated disorder arising because of the pres- ence of defects. According to Ref. 5, an accomodation line separates two regimes of behavior of the vortex matter: the regime of individual pinning, and the regime of collective pinning with the participation of linear defects.11 The posi- tion of the accomodation line is determined by the competi- tion and compensation of the characteristic elastic energy arising because of the interaction with the remaining vortex lattice ͑the potential box͒ and the pinning energy. Below the accomodation line the pinning force dominates, and the vor- tices can therefore be shifted substantially from their equilib- FIG. 2. Vortex line in a potential ‘‘box’’ containing an extended linear defect,5,10,11 which modulates the parabolic potential well, creating new rium positions in the perfect lattice. To emphasize the par- minima and displacing the vortex from its equilibrium position in the box. ticularly strong individual pinning in such a state, it has been Also shown are the short-wavelength lateral displacements that arise due to proposed5 to call this state of the vortex lattice a ‘‘strong the simultaneous action of a random point potential. These displacements Bose glass.’’ It was proposed5 that the state that is formed in promote depinning of the vortex from the linear defect. the collective pinning regime be considered a quasilattice or ‘‘weak Bose glass.’’A very important distinction between the tance for the controlled creation of desired transport current- effect of linear and point defects is that, unlike point defects, carrying properties of a material. The competition of these ͑ ͒ which stimulate lateral bending deformations of the vorti- two types of disorder was first studied experimentally in ces, linear pinning centers stabilize the residence of a given Refs. 12–18 and theoretically in Ref. 19. In particular, it was vortex line in the potential box formed by the elastic inter- shown in Ref. 19 that in the real cases the influence of the action with the neighbors against thermal fluctuations and correlated disorder is dominant in the behavior of the vortex ͑ against bending displacements caused by point defects as is matter, even when the density of defects is lower than the ͒ shown schematically in Fig. 2 , thereby preventing the for- density of vortices. mation of a three-dimensionally disordered solid state of the Up till now, however, the discussions and calculations of vortex lattice. It is particularly important to emphasize that in theorists have been based exclusively on the hypotheses of the case of linear crystal defects—pinning centers—the structures with linear defects obtained in HTSCs subjected to ‘‘strong’’ glass phase, with a highly two-dimensional dis- irradiation by heavy ions with high energies ͑around 1 GeV torted structure of the vortex lattice and the highest critical and higher͒. Very recently it has become clear that disloca- current density, occupies the low-field part of the H – T phase tion ensembles in strongly biaxially textured epitaxial films diagram. Of course, such behavior contrasts with the effects of YBCO play the role of strongly-pinning correlated linear of point disorder, when the more strongly pinned three- defects. dimensionally disordered solid vortex phase appears in higher fields.12–18 4. PINNING CENTERS IN EPITAXIAL YBCO FILMS Thus it turns out that the contributions of linear and point defects to the volume pinning force not only do not add As we have said, the fact that the highest Jc together but they may even subtract, i.e., the point disorder (у1010 A/m2 at 1 T and 77 K͒ can be obtained relatively promotes the depinning ͑breakaway͒ of vortex lines from easily in YBCO epitaxial films while remaining unachievable linear defects. This conjecture has been shown for any YBCO bulk materials should be attributed to features experimentally12–18 to be well confirmed for configurations of the defect structure of the films and, accordingly, to the in which the linear defects are oriented parallel to the mag- formation of a volume pinning force in them due to the most netic field ͑or, more precisely, to the direction of the vortex effective interaction, primarily with linear defects. Many at- lines͒. Since all real HTSC materials unavoidably contain tempts have been made to establish a linkage of the pinning pointlike defects ͑e.g., oxygen vacancies͒, the question of the force to the the screw dislocations that initiate the corre- dominant pinning mechanism is one of paramount impor- sponding three-dimensional mode of film growth during Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan 735 deposition ͑by the appearance and propagation of screw dislocations͒—the so-called ‘‘screw-mediated’’ growth20,21 —and to the growth steps on the surface of the film in the two-dimensional, ‘‘layer-by-layer,’’ growth mode. However, these attempts have been futile: for example, one can prepare films in which the concentration of screw dislocations will differ by an order of magnitude or more ͑from 5ϫ107 to 109 cmϪ2͒ while the critical current density will be the same, and vice versa. Moreover, the cores of the growth screw dislocations in YBCO films have been shown by electron microscope studies to have a characteristic transverse size that is much greater than the coherence length.22 Therefore, they can hardly be effective pinning centers. The experimen- tal evidence13,14,23–25 of a very high density of edge disloca- tions in YBCO epitaxial films is also not very convincing, since the investigators could scarcely believe that a disloca- tion line ͑i.e., the core of a dislocation͒ could in itself serve as a strong pinning center providing effective ‘‘core pin- ning.’’ Indeed, on the one hand, back in the early papers of Dew-Hughs26 it was shown that dislocations pin only when they comprise an ensemble which forms ‘‘walls’’ of a cellu- lar structure in metallic bcc alloys owing to the long-range ␦ ␦ 6 stress fields, which can lead to both l and Tc pinning. As to an isolated linear defect, in the papers by Dew-Hughes and his followers it was clearly stated that ‘‘isolated disloca- ␬ ͑ tions lead to a negligibly small change in GL’’ the param- eter of the Ginzburg–Landau theory͒. And, on the other hand, it is known from the theory of dislocations that the transverse cross section of the core can have an area of the order of the square of the Burgers vector, which is clearly much smaller than the square of the coherence length, even ͑␰ Ϸ for HTSCs ab 1.5 nm for YBCO in comparison with the value of the Burgers vector BϷ0.4 nm͒. Thus it appears that, as in the case of point defects, dislocations can pin only collectively, since the depth of the pinning potential well U p is small, and the experimental pinning force of an individual ϭ dislocation, f p dUp /dr, is small. Two new important re- sults have played a most important role in the further devel- opment of the concepts of dislocation pinning in HTSCs, presenting unambiguous evidence in favor of strong pinning on dislocation lines. The first of these was obtained by Chisholm and Smith,27 and then, more convincingly, by Merkle and co-workers.28,29 A high-resolution transmission FIG. 3. a—Moire´ pattern showing edge dislocations with the aid of a trans- electron microscope ͑HRTEM͒ was used to obtain pictures of mission electron microscope ͑TEM͒; the dislocation line are perpendicular to the cuprate planes. The dislocations, which are additional crystallographic the real distribution of atoms inside and in a neighborhood half planes in the structure are clearly revealed by the moire´ fringes, the around the core of a complete ͓001͔ edge dislocation. It was distance between which is 2.2–2.3 nm.13,14,18,25,35 b—TEM moire´ pattern found that the core of an edge dislocation is a highly dis- showing the distribution of edge dislocations ͑in a YBCO film 10–12 nm ͒ torted structure on a characteristic scale length of the order of thick, deposited on a single-crystal MgO substrate ; the dislocation lines are 28,29 perpendicular to the ab plane. The arrangement of the dislocations corre- 5–10 interatomic distances, i.e., around 3 nm. In addi- sponds to an averaged misorientation angle of 1.2° between neighboring tion, it was found that in the core region there is an excess of domains ͑subgrains͒.13,14,18,25,35 copper above stoichiometry ͑see, e.g., Ref. 28͒. Thus it turns out that the cores of dislocation lines have a structure that resembles that of artificial defects introduced by irradiation that our recent HRTEM results ͑some of which are shown in with heavy ions of high energy; those defects act as exceed- Fig. 3͒ unambiguously confirm the formation of such a dis- ingly strong one-dimensional linear pinning centers. This has location structure in YBCO films exhibiting record values of ͑ been shown convincingly in many studies see, e.g., Refs. Jc(H). These results also make it possible to understand the 30–32͒. mechanism of formation of linear defects during deposition It remained to show whether a structure state providing a and growth of the films. The problem of the formation of a very high average dislocation density is formed in YBCO dislocation substructure during epitaxial growth of films in epitaxial films during their deposition and growth. It appears general and YBCO films in particular goes beyond the scope 736 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan

of the present paper, and for an introduction to this topic the ideas.32,38,39 The model is actually completely adequate for reader is referred to the recent papers cited as Refs. 13, 14, treating the electrodynamics of bicrystals with this sort of tilt 18, 20–25, 27–29, 33, and 34. boundaries and with small values of the misorientation angle. The model does not require making any sort of assumptions 5. DYNAMICS OF VORTICES IN A YBCO as to the transmissivity for supercurrent or to Josephson SUPERCONDUCTOR WITH LINEAR PINNING CENTERS properties of the bicrystals with such a tilt dislocation bound- ary, as must be made for the models proposed in Refs. 40– It follows from the results of structural 13,14,18,20–25,27–29 43. Most importantly, in the framework of the model one can studies that the most interesting type of de- consistently explain the strong dependence of the critical cur- fects in YBCO thin epitaxial films and bicrystals in respect to rent J through the boundary on the misorientation angle ␽ their contribution to the pinning and dynamics of vortices are c between the adjacent domains ͑‘‘banks’’ of the junction͒,as low-angle tilt dislocation boundaries. Of course, point de- has been observed for YBCO bicrystals and films.35–37,44 In fects ͑oxygen vacancies͒ distributed in a random manner, order to apply this model for describing polydomain mosaic dislocation loops due to stacking faults and lying parallel to films containing this sort of low-angle boundaries, one must the ab plane, and ͕110͖ microtwins are present in the films, take into account the additional effects in the motion of vor- and in certain cases make an appreciable contribution to the tices in both the direction transverse to the domains and behavior of the vortices and to the magnetic-field depen- along the curved domain walls, with spatial variation of the dence of the critical current density. misorientation and also of the distance between adjacent dis- As we have said, recent transport measurements in com- locations. bination with a scaling analysis provide substantial argu- The model of vortex transport along a row of edge dis- ments in favor of the existence of a Bose glass phase in YBCO crystals and thin films irradiated by heavy ions at locations lying parallel to the c axis and forming, as we have high energy ͑1 GeV and higher͒. The formation of a Bose said, a low-angle boundary between slightly misoriented ͑ ͒ glass phase is initiated by randomly distributed correlated subgrains domains is based on the above treatment of the extended defects—columnar tracks made by the heavy ions. pinning and dynamics of vortices in superconductors with ͑ ͒ 11,32,35–38 However, as we have seen above, correlated linear defects in extended linear columnar defects. It is important deposited YBCO thin films are actually distributed nonuni- to note that collective effects were not taken into account, at formly: they form more or less well-formed rows or walls of least not in Refs. 35–37. According to the results of those edge dislocation lines lying parallel to one another, with non- papers, the depinning of vortices residing at linear pinning Ϸ␰ centers and their subsequent dynamics in the presence of superconducting cores of transverse size D ab . The dislo- cation walls frame domains of mosaicity, azimuthally misori- transport current occur owing to the spontaneous formation ented by approximately 1° with respect to the neighboring of vortex excitations caused by thermal fluctuations in the domains. Inside the domains the dominant pinning defects volume of the superconductor. A vortex excitation has the are pointlike ͑oxygen vacancies͒. For such a spatial distribu- form of a partially depinned vortex loop ͑Fig. 2͒. When the tion of correlated linear pinning centers the Bose glass phase size of the depinned part of such a loop exceeds a certain exists only at sufficiently low applied magnetic fields, for critical value for a given superconducting transport current ϰ ⌽ 1/2 ͑see Refs. 35 and 36͒, the loop becomes unstable and begins which a0 ( 0 /H) remains much larger than the trans- to expand until it touches the neighboring linear defect. After verse domain size Ldomain . When the magnetic field is in- that, the motion of the remaining part of the given vortex creased to a certain critical value HB→F there is a crossover to a distinctive new correlated glassy state, the ‘‘Fermi will occur through the motion of two vortex ‘‘steps’’ ͑kinks͒ glass,’’ in which the vortices are located both in dislocation moving apart in opposite directions along the c axis under walls and in the interior of the domains. The vortices inside the influence of the Lorentz force. The model of vortex mo- the domains interact with random point defects, and their tion along an equidistant row of mutually parallel disloca- behavior is therefore similar to that in single crystals. tions can be extended to the case of a real polydomain ͑mo- At a certain value of the applied magnetic field the linear saic͒ structure of a film with a low-angle-boundary network tension of the vortex ‘‘softens’’ as a result of dispersion of formed by ensembles of edge dislocations, as appears to be the elastic constant C44(k), and then the picture arises that is the case in an actual image similar to that shown in Fig. 3. described below in an analysis of the concrete results of our These dislocations are found in a more or less disordered ␪ recent experimental measurements of Jc(H,T, ) for YBCO state and are not equidistant, as was proposed above. Never- films ͑␪ is the angle between the cuprate planes and the ap- theless, using a percolation approach, as was done by plied magnetic field H͒. For an idealized case, i.e., when the Gurevich in a treatment of flux creep,45,46 one expects that low-angle boundary is formed by a regular row of equally the low-angle boundaries will serve as percolation channels spaced edge dislocations parallel to one another and with no for the thermally activated motion of vortices. This is appar- other disruptions of the crystalline order in neighboring sub- ently the case, since the activation energy Uc(D) for the grains ͑domains͒, one can use the model of vortex dynamics. transition of a vortex from one edge dislocation to a neigh- For this case Kasatkin35–37 examined the transport of vorti- boring one ͓i.e., at a distance D(␽)͔ is much less than the ces along low-angle domain walls ͑or more precisely, bicrys- corresponding value Uc0 for the traversal of a vortex across tal boundaries͒. In the framework of this model it was shown a domain. Thus the network of low-angle boundaries can be that despite the strong single-particle core pinning by lines of formed by a spatial landscape of Uc , in the more or less edge dislocations, the motion of the vortices occurs along the pronounced ‘‘valleys’’ of the vortex activation energy Uc(r). domain walls in accordance with the conventional Considering the thermally activated flux flow ͑TAFF͒ regime Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan 737

FIG. 5. Damping curves for Iϭ0 and 100 mA at Hϭ2 T. The dotted and dashed curves were calculated according to Eq. ͑3͒, following the procedure described in Refs. 48, 52, and 58.

YBCO films in the region of their depinning at angles ␪ϭ90° ͑which is equivalent to a parallel orientation of the applied field relative to the cuprate planes͒, 75°, and 60°, with the transport current always flowing perpendicular to the applied field. The dynamic behavior of the vortices in such an oscilla- FIG. 4. Square of the resonance frequency ͑a͒, of the corresponding damp- tor is well described by a diffusion model for the thermally ing ͑b͒, and of the resistivity ͑c͒ as functions of temperature for different activated vortices.50 In this model the diffusion constant is ϭ ␪ϭ values of the transport current at a field H 2 T and 0°. The dashed given by curve in part ͑c͒ is an example of a resistivity curve rescaled by means of Eq. ͑2͒ to the diffusion constant, which was taken for calculation of the DϭD ͒ϭ␳ ͒ ␮ ͑ ͒ ͑H,T ͑H,T / 0 , 2 damping curve ͑see Fig. 5͒ according to Eq. ͑3͒. where ␳ is the resistivity of flux flow. In the framework of this model, in accordance with Refs. 48, 51, and 52, the of vortex motion along these percolation channels, one ex- damping of the vibrating superconductor has the form pects that the average distance between nearest dislocations ␮ 2 1 0H is nevertheless related to the angle of misorientation of ⌫ϭ ͩ ͪ V ␹Љ͑D͒, ͑3͒ 2I ␻ 2 p neighboring domains. Then at small values of the angle one i ␽ ϰ ␽ Ϫ1 has ͗Uc(D)͘ ͗(sin ) ͘, and, hence, the strong depen- where Ii is the effective moment of inertia of the oscillator, ␽ 48 ␻ dence Jc( ) obtained for a straight row of equidistant dislo- which can be found experimentally, is the resonance fre- cations will hold for the percolation situation with sin ␽ re- quency of the oscillator, V p is the volume of the sample, and placed by its value averaged over the whole film. ␹Љ is the imaginary part of the ac susceptibility of the ␽ ␹Љ A strong dependence Jc( ) for polydomain epitaxial sample: as a function of temperature has a maximum ␹Љ ϭ 51 films of YBCO with low-angle boundaries was observed ex- max 0.41723. This result is valid for any value of the ␹Љ perimentally in Ref. 47. transport current density J, i.e., max is independent of the current load. Changing the diffusion parameter D by the pas- 5.1. Dynamic depinning of the vortex lattice in YBCO films sage of current only shifts the position of the damping peak An extremely informative method for describing the be- ͑e.g., on account of the dependence of the effective activa- ͒ havior of the vortex ensemble in YBCO films in the dynamic tion barrier U p on the current , but it does not introduce any regime utilizes a mechanical oscillator with a high Q factor changes in its absolute value. It should be noted that the top under passage of a dc transport current. Such measurements of the damping peak determines the position of the depinning have been done using miniature current and potential con- temperature Tdp or the irreversibility point. tacts on a vibrating YBCO film sample.48 The technique per- If the measured value of the resistivity ␳(T,H,J) is used ͒ Dϭ␳ ␮ mits making the following simultaneous measurements: 1 for / 0 according to the procedure described in Ref. the change in the square of the resonance frequency, 48, then it becomes possible to apply the theory quantita- ␻2 Ϫ␻2 (Ba ,T) (0,T), of a mechanical oscillator with a film tively to the discussion of the experimental results, through sample of the HTSC YBCO attached to it, executing vibra- the use of Eq. ͑3͒. Figure 5 shows the the calculated damping tional motion in a magnetic field, which yields information curves for two values of the current Iϭ0mA(Jϭ0 A/m2) about the pinning force49,50 ͑Fig. 4a͒;2͒ the damping ⌫, and Iϭ100 mA (JϷ2ϫ108 A/m2). The calculated tempera- which characterizes dissipative processes in the sample in ture dependence for ⌫(Iϭ100 mA), its absolute value, and connection with the motion of the vortices49,50 ͑Fig. 4b͒;3͒ the width of the transition turn out to be in good agreement the resistivity of YBCO films ͑Fig. 4c͒. Using this technique with experiment. Furthermore, it is clear that richer informa- we were able to study the dynamic behavior of vortices in tion about the depinning transition of the vortex lattice can 738 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan be obtained from the damping of a vibrating superconductor than from the value of the electric field induced by the mo- tion of the vortex lattice. The broadening and asymmetry of the damping peak are evidence of a diverse character of the depinning of the vortex lattice. Of course, from the present investigations it is difficult to reconstruct the microscopic picture of the dynamics of the moving vortices, but, considering the known results from observations of the motion of an almost ordered vortex lat- 53 FIG. 6. Pinning potential with a vibrating vortex at different temperatures tice in superconducting crystals, the results of theoretical Ͻ Ͻ (T1 T2 T3) during the passage of current and with no current. papers on moving disordered vortex lattices,54,55 and the aforementioned complex ‘‘valley’’ spatial landscape of the pinning potential U in real YBCO superconducting films, p valleys,’’ i.e., the easy-slip channels, when they are oriented we propose a scenario that will lead to the observed effects parallel to the driving force ͑see also Ref. 60͒. This behavior, ͑Fig. 4b͒:1͒ a narrowing and symmetrizing of the damping as we have said, can be treated as percolational.45,46 If the peak upon application of a current load; 2͒ a rise of the peak initial breakaway of the vortices when depinning is achieved value of the damping for JϾ2ϫ106 A/m2. We note that at is interpreted as vortex rows moving along ‘‘channels,’’ then high values of the transport current density the height of the one expects that the vortex lattice inside the domains remains damping peak becomes independent of the current density, pinned on account of the strong pinning on domain walls and only the position of the peak changes, shifting to lower perpendicular to the driving force, i.e., it remains practically temperatures with increasing current. Ͻ immobile at temperatures T Tdp . As the temperature is In the YBCO films under study there is a chaotic distri- ϭ raised, the observed damping peak at Tdp(J 0) is indicative bution of the pinning potential, the interaction with which of depinning of strongly pinned regions. Under the influence leads to disordering of the vortex lattice. The situation be- of the aforementioned factors, this peak, which characterizes come even more complicated when one takes into account the depinning of the vortex lattice, becomes broad and asym- that rather complex vortex configurations can arise for field metric. orientations close to parallel with respect to the ab planes, Phenomenologically, without specifying the type of pin- when two mutually perpendicular vortex subsystems can co- ning defects, one can describe the behavior of the damping exist simultaneously in the superconducting films on account peak of a vibrating sample at Jϭ0 by introducing a distri- of the two components of the field vector penetrating into the bution of the potential energy of the activation barriers. Such 56–58 film. Then for vortices parallel to the ab plane the ma- a distribution will lead to two main effects: broadening of the jority of linear defects are dislocation loops distributed cha- transition with respect to temperature, and, consequently, de- otically in the interior of the film. Their characteristic size is creasing the height of the damping peak at the depinning 5–7 nm, but then their pinning force hardly decreases with temperature. Indeed, the use of a single-barrier approxima- temperature because of the weak stress–strain fields near the tion in the framework of the diffusion model leads to an core ͑see Sec. 7͒ and also because in this orientation the unavoidable difference between theory and experiment, as is 48,52 bending modulus C44 does not ‘‘soften’’ along the field, as it usually observed in practice. does in the case Hʈc on account of the anisotropy and lay- One can give a simple picture in which the loading of ered nature of the atomic structure. In the case of vortices the sample by current leads to tilting of the whole pinning parallel to the c axis there exists a hierarchy of pinning cen- potential61 and in which the depth of the potential is tempera- ters, with a spatial distribution: 1͒ the strongest obstacles for ture dependent ͑Fig. 6͒. By scanning the pinning potential the transverse motion of the vortices are close-packed dislo- well with the aid of low-amplitude mechanical vibrations of cation walls with vortices already residing on them and with the superconducting sample, one can record the tiniest shielding currents flowing along these walls; 2͒ individual changes in the height of this well at very low transport cur- ‘‘standing’’ or ‘‘threading’’ dislocations, which are also rents. This effect is observed at current densities as low as strongly pinning if Hʈc;3͒ point defects inside the domains, JϽ106 A/m2. The decrease and shift of the high-temperature which can rapidly decrease the net pinning force and current part of the transition ͑Fig. 4͒ indicates that the upper part of ͑ ͑ ͒ density, as follows from the arguments given above see Sec. the potential distribution high U p is shifted to lower values ͒ ͑ ͒ 3 and the experimental data presented below Sec. 8 . of U p . It is possible that a decrease and shift to higher values ͑ The larger-scale topological defects existing in YBCO of U p occurs in the low-lying part of the U p distribution i.e., films, such as shaped ͑polygonized͒ domain dislocation at lower temperatures͒. This effect can scarcely be observed walls, can lead to plastic flow of vortex rows but not to by the technique used, since the flux bundles remain pinned displacement of an elastically deformable vortex lattice. A by the highest potential barriers until the temperature be- mechanism for the motion of vortices along so-called ‘‘easy- comes high enough to provoke depinning. slip channels’’ has been proposed previously ͑see, e.g., Ref. At large values of the driving force (Jӷ106 A/m2) the 59͒. However, this mechanism requires a small shear modu- velocity of the vortices is expected to play the role of an lus C66 in the vortex lattice, and that ordinarily takes place in order parameter. The observed motional narrowing of the fields approaching Hirr . Thus the thermally activated depin- damping peak upon increasing current loading ͑i.e., driving ning of the vortex lattice under the influence of a small driv- force͒ means that a new ‘‘instrument’’ for the ordering of the ing force can occur near Hirr(T) along ‘‘plastic flow vortex lattice has appeared. We again note the explicit satu- Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan 739

FIG. 7. Normalized damping peak ͑a͒ and the half-width of the damping FIG. 8. a—Resistivity of a YBCO film versus the transport current density, peak ͑b͒ of a vibrating YBCO film as functions of the transport current measured at Hϭ2 T and various temperatures with a step of 0.1 K. The density at different values of the external magnetic field. The dashed curves dotted line is the so-called ‘‘vortex glass’’ transition at a temperature Tg were calculated with the use of the function aϩbtanh(J/c), where aϭ0.57 ϭ86.35 K. b—Scaling behavior of the ␳(J) curves as described in the text; 6 6 2 ͑1.03͒, bϭ0.42 ͑0.44͒, and ͉c͉ϭ15ϫ10 (5.5ϫ10 )A/m for the upper here zϷ1.7 and ␣ϭ4.7Ϯ1. ͑lower͒ panel.

The ‘‘vortex glass’’ model predicts that the typical scal- ration effect at high values of the transport current ͑Fig. 7͒. ing procedure can be applied to the ␳(J) curves to give two In terms of the picture of a distribution of effective potential groups of lines with different signs of the curvature. Indeed, barriers, this result indicates that the moving elastic vortex the experimental curves can be subjected to a scaling analy- ␳ ͉ Ϫ ͉␣(zϪ1) ͉ lattice becomes increasingly dominant over the pinning dis- sis in the coordinates ab / T Tg versus J/T T Ϫ ͉2␣ ϭ Ϯ ␣ϭ Ϯ ͑ ͒ order and approaches a delta-function distribution of barri- Tg , where z 1.7 0.15 and 4.7 1 Fig. 8b . Both ers. In other words, the thermally activated depinning be- parameters agree with the values given in Refs. 63 and 64. haves as if all the barriers have equal depth. It can also be A particularly interesting fact that emerges from these assumed that the vortex lattice begins to behave like an elas- measurements is that for Iϭ0 and angles ␪ϭ90°, 75°, and tically deformable continuum whose motion is insensitive to 60° in magnetic fields of 0.5 TрHр2 T the depinning tem- the individual pinning centers. An important circumstance is perature and the temperature of the proposed formation of ⌫ Х that the damping (T) of the vibrating film can be described the ‘‘vortex glass’’ coincide: Tg Tdp . Since the behavior quantitatively with the help of formula ͑3͒ without any free observed with the use of the mechanical oscillator can be parameters by using the experimentally determined resistiv- described by a depinning transition in the framework of dif- ity of the flux flow as the diffusion constant ͑2͒. fusion concepts, it is necessary to understand to what extent the result obtained from the scaling analysis can be taken as evidence of a second-order phase transition in the vortex 5.2. Vortex glass or depinning? lattice. In contrast to the experimentally proven melting of Simultaneous measurements of the resonance frequency the vortex lattice, which has been observed in clean high- and damping of a vibrating sample of a superconducting temperature superconductors by different techniques, the YBCO film together with the resistivity curves ␳(T,J) ͑Fig. ‘‘vortex glass’’ transition is determined solely by means of 4͒ permit a direct comparison of the so-called ‘‘glass tem- the scaling analysis shown in Fig. 8b. One wonders whether perature’’ Tg and the depinning temperature Tdp . Figure 8a the result of this scaling analysis can be described using a shows the dependence of the resistivity on the transport cur- different model without invoking the concepts of a ‘‘vortex rent density. It is seen that the ␳(J) curves are clearly sepa- glass’’ and a phase transition of the vortex lattice into such a ϭ rated into two groups by the dotted line Tg 86.35 K. Be- state. havior of this type is often used as an argument in favor of Indeed, a serious alternative interpretation of the results the possibility of a ‘‘vortex glass’’ state.7,8,62 In other words, of the scaling analysis of the experimental data is the perco- it is assumed that the vortex lattice undergoes a second-order lation model of vortex motion proposed by a number of phase transition to a ‘‘vortex glass’’ state at a temperature authors over the past few years.65–67 According to the perco- Tg . lation model, a superconductor found in the mixed state con- 740 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan

FIG. 9. a—Angle dependence of Jc(H) measured by the four-probe transport method for YBCO films deposited by a pulsed laser technique on a LaAlO3 substrate. b—Schematic representation of the angle dependence of Jc(H) for the case when both dislocation ensembles perpendicular and parallel to the ab plane contribute to the two-peak curves: the contributions from linear defects ͑I͒, from the intersection of vortices with dislocations ͑II͒, from pinning centers ͑ ͒ ͑ ͒ lying in the ab plane III , and from the anisotropy of Hirr IV .

sists of percolation regions with different resistivities, which relaxation time of the vortex lattice ͑the increasing velocity correspond to pinned and free ͑depinned͒ vortex lines. When the lattice becomes ‘‘stiffer’’͒. On the whole, this process the percolation threshold is reached there is a transition from leads to an average viscosity ␩ of the vortex lattice which is Dϭ␮ 2 ␩ the depinned state to a pinned state with an insignificant related to the diffusion constant as 0H / . linear resistivity. In the framework of the percolation picture the disorder and the distribution of percolation regions also leads to a broader pinning transition. Consequently, the nar- ␪ 6. THE Jc„H,T, … CURVES FOR YBCO FILMS. RELATION rowing of the depinning transition when a current load is TO THE CHARACTERISTIC ENSEMBLE OF LINEAR applied, as we have said, can be interpreted as a decrease of DEFECTS the disorder ͑or the pinning͒ of the vortex lattice as a result of its motion. In the ‘‘microscopic’’ picture of this process, Saemann-Ischenko68 was the first to show that when the the vortex lattice, as its velocity increases, ceases to ‘‘notice’’ magnetic field vector is rotated with respect to the c axis the point and pointlike defects ͑e.g., the crossing of disloca- ͑while maintaining constancy of the Lorentz force͒ in a bi- ͒ ␪ tion lines by vortices , through which a vortex passes in a axially textured epitaxial YBCO film the Jc(H,T, ) curves very short time. There is insufficient time for the vortex to have two characteristic peaks at Hʈc (␪ϭ0°) and Hʈab interact with them, since the pinning time is shorter than the (␪ϭ90°), which are shown in Fig. 9 ͑our data͒. The prop-

␪ TABLE I. Properties, characteristics, and behavior of the Jc( ) peaks.

␪ ʈ ␪ ʈ Experimentally measured parameter Jc( ) peak for H cJc( ) peak for H ab 1. Temperature change during measurements Relative height grows with Relative height decreases with → increasing T for 40–80 K, then the peak is increasing T, but for T Tirr(H)the → suppressed for T Tirr(H). peak becomes dominant. 2. Influence of the strength of the applied Practically vanishes in fields above The higher the field, the sharper the peak. → magnetic field 2–3 T. For T Tirr(H) only this peak remains. 3. Rate and method of deposition of the film Low-rate off-axis magnetron High-rate pulsed laser deposition sputtering tends to enhance substantially enhances this peak. the peak ͑e.g., for ͒ sapphire/CeO2 /YBCO films . 4. Effects of the substrate and buffer layers The greater the misfit of the crystal This peak is sharper and more significant lattices between the materials if the misfit is small, of the substrate, buffer layers͑s͒, as, e.g., for substrates of

and YBCO film, the stronger the peak. SrTiO3 or LaAlO3. 5. Influence of deposition temperature Height increases with increasing Substantially suppressed with increasing of YBCO film T ͑up to 745–750 °C͒ for T ͑up to 745–750 °C͒ in the case of

sapphire/CeO2 /YBCO films. sapphire/CeO2/BCO films. 6. Effect of a change in thickness of the The thicker the YBCO film ͑at least Practically independent of the thickness of ͒ YBCO film on a LaAlO3 substrate , the higher the peak. the YBCO film. 7. Effect of a change in the growth mode of the 2D growth mode suppresses the peak and 3D 3D growth mode leads to substantial YBCO film mode enhances it. suppression of the peak. Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan 741 erties, characteristics, and behavior of the two peaks are de- scribed below and listed in Table I. It can be seen from Table I that the relative height of these peaks depends on the conditions of measurement of the current–voltage characteristics and determination of the value of the critical current, specifically, on the temperature of the measurements, the applied magnetic field, and the value of the electric field Ec . Indeed, the Jc(H) peak for Hʈc initially grows with increasing temperature approxi- mately from 40 to 80 K and then, when the temperature approaches the line of irreversibility, the Jc peak is substan- tially suppressed and can even vanish in the case of YBCO films deposited by pulsed laser sputtering on LaAlO3 sub- ʈ strates. On the contrary, the Jc(H) peak for H ab survives at all temperatures, magnetic fields, and velocities of the vortex lattice under the influence of the Lorentz force. As is now known from the high-resolution electron mi- croscopy data, there are several types of dislocation en- sembles that can form in YBCO films during their growth: 1. Misfit edge dislocations at the boundary due to the usual mismatch of the interatomic spacings in the crystal FIG. 10. Schematic illustrations of ensembles of edge dislocations, which lattices of the substrate and growing film. are formed in YBCO films during growth in different modes: a—ensembles of ‘‘threading’’ dislocations perpendicular to the cuprate planes and lying in 2. Dislocation loops due to the existence of stacking low-angle tilt domain boundaries ͑substrate with large misfit ͑MgO, YSZ, ͑ ϩ ͒͒ faults i.e., the local appearance of ‘‘extra’’ or ‘‘missing’’ sapphire CeO2 ; b—ensembles of dislocations parallel to the cuprate segments of CuO2 planes, usually up to 10 nm in size, during planes and consisting of misfit dislocations at the boundaries and dislocation ͑ ͑ growth͒; these are edge dislocations, the dislocation lines loops induced by local stacking faults substrate with small misfit SrTiO3, LaAlO ͒͒. being parallel to the ab plane.69 3 3. ‘‘Threading’’ edge dislocations, whose dislocation lines, being parallel to the c axis and perpendicular to the lying in the ab plane. It is important to note that these latter surface of the film, as a result of the polygonization process dislocations and loops also contribute to the maximum partially or completely form low-angle tilt boundaries of do- ʈ Jc(H c), since, as was first shown in Ref. 73 and then con- mains of azimuthal mosaicity in the film. The average den- firmed by the present authors in Refs. 12–14, 74, and 75, 11 2 sity of such dislocation lines can reach 10 lines/cm and when the vortices are parallel to the c axis they can interact 13,14,18,23–25,31 even higher. These dislocations are formed with transverse dislocation lines as with pointlike defects. mainly as a result of the realization of a two-dimensional ␪ Then the function Jc(H, ) has a domelike character of the heteroepitaxial growth mode in which a so-called ‘‘rota- ␪ ϰ ␪ 1/2 type Jc( ) (cos ) . It can also be assumed that some con- tional’’ relaxation of the misfit at the boundary occurs. In tribution to the volume pinning force is made by point de- particular, this mechanism can be enhanced further on ac- ʈ fects such as oxygen vacancies. At the peak Jc(H c) the two count of specific growth conditions, e.g., as a consequence of contributions from point and quasi-point pinning centers cre- their rotational misfit on the R plane of sapphire and the ate the pedestal that vanishes as the line of irreversibility is ͑ ͒ ͑ ͒ → 001 plane of CeO2 for details see, e.g., Refs. 70 and 71 . approached, i.e., for T Tirr(H), as the magnetic field is 4. Screw dislocations at the boundary in low-angle twist increased, and as the velocity of the vortices ͑i.e., the dimen- ͒ 12–14,18 boundaries, which are a source of screw dislocations in the sionless number Ec increases. As to the peak ʈ low-angle domain walls, making them increasingly more Jc(H ab), since it survives at all temperatures, fields, and complex tilt–rotational boundaries the higher the degree of velocities of the vortex lattice, we may assume that it is due mismatch of the crystal lattices at the boundary. Screw dis- to dislocation pinning on dislocations and loops lying in the location sources emerging on the surface of the film at higher ab plane and also to anisotropy of Hirr(T). As was discussed ͑ ͒ deposition temperatures above 740 °C for YBCO can ini- above, anisotropy of Hirr(T) leads to anisotropy of the resis- tiate a three-dimensional growth mechanism with the forma- tivity due to the motion of the vortex lattice under the influ- tion of polygonal spirals.20,25,35,72 ence of the Lorentz force. This means that at a given value of Analysis of the data obtained as a result of measure- Ec the value of Jc determined from the current–voltage char- ʈ ʈ ments of the angle dependence of Jc(H) for a large number acteristics for H ab and H c will be substantially different. of perfect biaxially textured YBCO films with high Jc(H) Thus, in considering the behavior of YBCO films one shows convincingly that the maxima of the critical current should keep in mind that there are two ensembles of edge and hence the orientation of the dislocation ensemble corre- dislocations: 1͒ with dislocation lines along the c axis, and 2͒ spond to two directions: parallel to the c axis, and parallel to in the ab plane. This was confirmed in our recent papers74–77 the ab plane. This conclusion is equivalent to the assumption by means of transmission and high-resolution electron mi- that the effective pinning of vortices in YBCO films is due to croscopy and also by electronic transport measurements of extended linear defects oriented along the given directions, the angle dependence of the critical current density in a mag- i.e., dislocation ensembles a ͓100͔ and dislocations and loops netic field. Such measurements have also been made by other 742 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan

an isotropically, hydrostatically compressed crystal is very weak. However, in regions of the crystal with a local aniso- tropic deformation the changes of Tc can be significant. In the linear approximation this dependence can be written as ͒ϭ Ϫ ␧ ͒ ͑ ͒ Tc͑r Tc0 Cik ik͑r . 4

Here Tc0 is the critical temperature of the undeformed crys- ␧ tal, ik is the strain tensor, and the coefficients ץ ץ ␧ץ ץϭϪ Cik Tc / ik are related to the derivatives Tc / Pa ͑Fig. 11͒. According to Refs. 79 and 81, the diagonal coeffi- cients Cii in the ab basal plane of the crystal are ␧ Ϸץ ץ␧ ϷϪ ϭץ ץϭϪ Ca Tc / aa 220 K, Cb Tc / bb 320 K, and ␧ Ϸץ ץ Tc / cc 0. Based on the experimental data,79–81 a theoretical calcu- lation by Gurevich and Pashitskii82 showed that the elastic deformations created by a single edge dislocation or a dislo- cation ensemble ͑e.g., a ‘‘wall’’ of dislocations͒ in an aniso- tropic crystal can cause a local elevation or depression of Tc FIG. 11. Schematic phase diagram of an anisotropic metal-oxide cuprate of and can even suppress the superconducting state completely the YBCO type with a high Tc . This diagram was obtained by Gurevich and ͑ ͒ 82 at a given temperature e.g., at 77 K . Consequently, the Pashitskii from the well-known Tc(x) phase diagram, where x is the dop- ͑ ͒ ing level, with the use of Eq. ͑4͒. region of the normal nonsuperconducting phase should ex- ist around a dislocation core, which, as described above in accordance with the data of Refs. 28 and 29, is ͑for a groups.68,78 For a clearer and more convincing demonstration ‘‘threading’’ edge dislocation a ͓100͔͒ a cylindrical channel of the real nonuniform distribution of dislocation ensembles of highly plastically deformed medium with a diameter of in YBCO films prepared under different conditions of nucle- around 2 nm. The normal regions surrounding the core ation and growth and by different growth mechanisms, in should also play an important role in the formation of the Fig. 10 we show schematic illustrations of the fine structure pinning potential of the superconductor. Therefore it is nec- of the films: with a set of mosaic domains differing from one essary to examine in more detail the deformation mechanism another by low-angle tilt dislocation boundaries ͑a typical of suppression of superconductivity in the neighborhood of a misorientation angle of around 1 – 2°, typical domain size of dislocation core, i.e., of the effect of an elastic strain field in 20–50 nm, average distance between dislocations in the the anisotropic crystal YBa2Cu3O7Ϫ␦ . boundary 10–20 nm, and average dislocation density around An edge dislocation perpendicular to the ab plane gives 1011 lines/cm2 ͑Fig. 10a͒; with misfit dislocations at the rise to elastic strains in the ab plane and to corresponding boundary and dislocation loops due to stacking faults, i.e., local changes in Tc . Under certain conditions, if the Burgers with extra or missing segments of the copper–oxygen layers vector B is directed at an angle ␽ to either the a or b axis, 69 of the CuO2 type, with the dislocation lines lying in the ab the change of Tc is given by the expression plane and an extremely high average dislocation density ␦T ϭϪC͓␧ϩ␤͑␧ Ϫ␧ ͒cos 2␽ϩ2␤␧ sin 2␽͔, ͑5͒ ͑Fig. 10b͒. It is surprising that in YBCO films such a dislo- c xx yy xy cation structure with a high density can coexist with a very where high degree of perfection of the crystal structure, as charac- ϩ Ϫ Ca Cb Ca Cb terized by transmission and high-resolution electron ␧ϭ␧ ϩ␧ ;Cϭ ;␤ϭ . ͑6͒ xx yy 2 C ϩC microscopy23–25,27–29 and x-ray diffractometry.74 However, a b 83 ␧ the most reliable tool for experimental observation of these Using the well-known components of the strain tensor ik two different dislocation ensembles is measurement of the under the condition Bʈa, Pashitskii74 obtained an expression ␪ angle dependence of Jc(H, ) at a constant value of the for the variation of the critical temperature in cylindrical co- Lorentz force74–77 ͑see Fig. 9͒. ordinates: CB sin ␸ 7. ELASTIC STRAIN FIELDS IN THE NEIGHBORHOOD ␦T ͑r,␸͒ϭϪ ͓͑1Ϫ2␴͒ϩ2␤ cos2 ␸͔. c 2␲͑1Ϫ␴͒ 2 OF LINEAR DEFECTS AND THE FEATURES OF THE PINNING ͑7͒ POTENTIAL IN YBCO FILMS Here ␸ is the azimuthal angle in the ab plane, measured It has been shown experimentally79–81 that layered an- from the a or b axis, ␴Ϸ0.28 is Poisson’s ratio,84 and B is isotropic HTSC metal-oxide cuprates have an anomalously the modulus of the Burgers vector, which is approximately strong anisotropic dependence of the critical temperature T c equal to the lattice constant aϷ0.4 nm in the ab plane. Thus on the pressure in the case of uniaxial compression. For 74 80 Pashitskii determined the boundary of the region of the example, for the optimally doped YBa2Cu3O7Ϫ␦ single normal phase around a dislocation core, i.e., the region in ץ ץ crystal the derivatives Tc / Pi measured along the principal ϷϪ which the local value of Tc is lower than the characteristic ץ ץ crystallographic axes are: Tc / Pa (1.9– 2) K/GPa, :ϷϪ average level ץ ץ Ϸ ץ ץ Tc / Pb (1.9– 2.2) K/GPa, and Tc / Pc (0 – 0.3) ␸ ͒ϭ ͒ ␸ ϩ␤ 2 ␸ у ͑ ͒ K/GPa. This means that the pressure dependence Tc(P) for rN͑ ,T R0͑T sin ͓1 0 cos ͔ 0, 8 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan 743

␪ FIG. 13. Temperature dependence of the height of the Jc(H, ) peaks for orientations of the magnetic field Hʈc and Hʈab, shown in relative units in FIG. 12. Shape and size of the region of the normal state at 77 K, and also relation to the position of the pedestal ͑or ‘‘background’’͒. the region partially suppressed superconducting order parameter in the neighborhood of a dislocation core for a ‘‘threading’’ dislocation whose dislocation line is perpendicular to the ab plane of the YBCO crystal; cal- culated by Pashitskii74 with the use of Eqs. ͑4͒, ͑7͒, and ͑8͒. allel to the ab plane, unlike the perpendicular ones, have insignificant strain fields in the neighborhood of the core. This means that the corresponding pinning potential wells are much narrower and steeper. The elementary pinning where force for these dislocations is larger. In addition, their shape CB͑1Ϫ2␴͒ T 2␤ is apparently weakly dependent on temperature, and hence R ͑T͒ϭ , ␶ϭ1Ϫ , ␤ ϭ . 0 2␲͑1Ϫ␴͒T ␶ T 0 1Ϫ2␴ the pinning force is independent of temperature, as is con- c0 c0 ͑ ͒ ͑9͒ firmed in experiment Fig. 13 . The opposite forms of the temperature dependence of the relative height of the J peaks ϭ c For Tc0 90 K and for the values of the coefficients Ca,b for Hʈab and Hʈc is evidently due in large measure to the Ϸ ␶ mentioned above, the parameter R0(T) 0.042/ (nm), and fact that the pinning potential wells have a different shape ␤ ϷϪ ϭ Ϸ 0 24. At T 77 K we have R0 0.3 nm and the maxi- and different temperature behavior: steeper slopes and weak ␸ϭϪ␲ mum of rN occurs in the direction /4 and has the temperature dependence for Hʈab, and more gradual slopes Ϸ ͑ ͒ value rN max 2.5 nm Fig. 12 . and strong temperature dependence for Hʈc. If now we take into account the proximity effect, then it becomes clear that the region with a suppressed supercon- 8. MAGNETIC FIELD DEPENDENCE OF J H AND THE ducting order parameter extends in all directions to a dis- c„ … ␰ ϭ␰ ␶1/2 H – T PHASE DIAGRAM FOR AN ANISOTROPIC HTSC WITH tance of the order of the coherence length (T) 0 / ¸ ͑ ␰ Ϸ ϭ ͒ LINEAR DEFECTS FOR H c where 0 1.3 nm is the coherence length at T 0 . ʈ At 77 K we have ␰(T)Ϸ3.5 nm, and the maximum The typical Jc(H) curves of YBCO films for H c,as ϭ ϩ␰ width of the normal region is L(T) 2͓rN max (T)͔ was shown in Refs. 74–77, 85, and 86 have three different у12 nm. The area of the region of suppressed order param- segments ͑Fig. 14͒. The low-field part is a plateau or a ϭ eter per dislocation parallel to the c axis can be estimated as weakly field-dependent part up to the point H HA , which Ϸ ␰ Ϸ ϫ Ϫ13 2 Ϸ ⌽ 1/2Ϸ␭ ͑ϳ ͒ SN(T) 2 (T)L(T) 8 10 cm at 77 K. This means corresponds to a0 ( 0 /HA) L 0.1Tat77K, that at a concentration of edge dislocations of around where the position of the points HA must be determined as 11 2 10 lines/cm the fraction of the normal phase is approxi- the intersection of the horizontal part of the curve Jc(H) Ϸ ϰ Ϫ0.5 mately equal to 10%. Each component of the function L(T) const with the extrapolated part Jc H , but on our → ϰ␶Ϫ1 ͑ ͒ increases at a different rate as T Tc0 : rN(T) , and curves Fig. 14 this cannot always be done because of the ␰(T)ϰ␶Ϫ1/2. Consequently, the width and shape of the pin- lack of measurements in sufficiently low fields. Undoubtedly ning potential well change with increasing temperature faster ␶Ϫ1 ␶Ϫ3/2 → than , the dependence approaching as T Tc0 . Al- though a detailed analysis of the consequences of this situa- tion goes beyond the scope of this paper, there are sufficient grounds for assuming that the pinning force arising in the interaction of vortices with dislocations of this ensemble should decrease appreciably with increasing temperature, as is observed in the measurement of the temperature depen- ʈ ͑ ͒ dence of Jc(H c) Fig. 13 . We conclude this Section by noting that the anisotropy ␤ ͑ ͒ parameter 0 in Eq. 8 for dislocation lines parallel to the ab plane is smaller by many times than that for perpendicu- lar ‘‘threading’’ dislocations, because of the neglibibly small ␤ϭ Ϫ ␴ Ϸ value of Cc . In this case 2/(1 2 ) 4.5. Thus one can FIG. 14. Magnetic-field dependence for biaxially oriented epitaxial YBCO assume that the dislocations and dislocation loops lying par- films for Hʈc. 744 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan here we are dealing with a regime of individual or single- particle pinning, in which the vortices are far apart and non- interacting, and a vortex lattice is not formed. This effect was also mentioned in Refs. 85 and 86, but the authors explained it as a transition through a so-called ‘‘matching’’ field HϭH⌽ , at which the density of vortices becomes equal to the density of linear pinning defects parallel to them. This is clearly incorrect for two reasons: first the authors of Refs. 85 and 86 incorrectly determined the density of defects from the etch pits, obtaining a value too low by one or two orders of magnitude ͑108 –109 lines/cm2 instead of the actual 1010–1011 lines/cm2, as is shown in Refs. 23–25͒; second, as the ‘‘matching’’ field is approached in YBCO, it is known87 ʈ that Jc(H c) increases rather than remaining constant. The intermediate part of the field dependence of ʈ Jc(H c) turns out to be practically linear on a log–log scale: FIG. 15. Proposed H – T phase diagram of the vortex matter in a YBCO Ϫ ϰ q ʈ Jc H , where the exponent q is close to 0.5. Such a de- epitaxial film for H c. pendence can be the result of the presence of a two- dimensional correlated ‘‘quenched’’ disorder in the vortex lattice, apparently of the Bose glass type, in this interval of This crossover in the behavior of the vortex system can be fields (H ϽHϽH ). At higher fields (HϾH ) the J (H) A B B c classified as a manifestation of a change in dimensionality in curve begins to decrease faster, viz., as HϪq with q the nonuniform ensemble. We denote the nonuniform state of ϭ1.0– 1.5. the vortex ensemble, interacting with the 2D pinning poten- A tentative explanation for this is that the vortex lattice tial parallel to equidistant linear centers forming a polygonal begins to interact with random pointlike defects as well. In- 2D lattice ͑if the field is applied parallel to the c axis͒,74–77 deed, at the crossover field HB the intervortex distance a0 as a correlated inhomogeneous Bose glass ͑CIBG͒. This state Ϸ ⌽ 1/2 ( 0 /HB) becomes comparable to the transverse size Ld differs substantially from the usual state of a triangular vor- of the domains of mosaicity, which are slightly misoriented tex lattice and a random Bose glass, in which the linear pin- with respect to one another and are separated by low-angle ning centers are distributed chaotically. The CIBG state dislocation boundaries. This is the most important point in ͑which is realized for H ϽHϽH ͒ also differs from the Ͻ Ͻ A B this treatment, since at low fields (HA H HB) the vortex Ͼ nonuniform state of the vortex lattice formed for H HB , lattice ‘‘perceives’’ the network of threading dislocations as a where the vortices interact with both linear and planar de- random, chaotic system which induces a quasi-two- fects. The latter can be called a Fermi glass, since the rela- dimensional correlated state of the vortex lattice of the Bose tion between the number of vortices resident on dislocations glass type with a high value of the bending modulus C44 . and the number of vortices inside the domains can be found When the field increases to HB , some of the vortices no by introducing a distribution of the Fermi type, much as was longer have a chance to become pinned on the dislocation done by Gurevich.45,46 Figure 15 shows the proposed H – T lines in the low-angle boundaries, since the benefit in terms phase diagram of the states of the vortex lattice in a YBCO of the pinning energy would be much smaller than the energy film for a magnetic field direction Hʈc. cost due to the 2D deformation of the vortex lattice. As a result, a significant number of vortices are inside the do- mains, being only weakly pinned by pointlike defects. Since CONCLUSIONS the magnetic field in this case is already quite high, one expects that a substantial change in behavior will occur, —Several type of dislocations and dislocation ensembles much as in the case of single crystals. Thus one expects an are formed in YBCO films during their growth: 1͒ edge dis- intradomain crossover from 2D to 3D behavior of the vortex locations of nonregistry at the boundary; 2͒ dislocation loops lattice due to wave-vector dispersion of the bending modulus due to stacking faults ͑i.e., with extra or missing segments of ͒ C44(k) of the vortex lattice, which is interacting with a ran- the CuO2 planes , the lines of which are parallel to the ab dom pointlike pinning potential. However, because of this plane; 3͒ edge dislocations in low-angle boundaries of mo- the phenomena that occur are radically different from that saic domains with a density of up to 1011 lines/cm2 and which occurs in a single crystal. In a single crystal the vor- higher. tices, by ‘‘softening,’’ more easily adjust to the chaotically —The efficient pinning of vortices and high critical cur- ͑ у ϫ 10 2 ͒ distributed point centers, and the resulting bulk pinning force rent densities Jc 3 10 A/m at 77 K in epitaxial YBCO 8 increases, and Jc(H) also increases from 10 to 5 films are due to the high density of linear defects formed in ϫ108 A/m2 ͑Refs. 12–18͒. In this case the vortices are the process of nucleation and growth of the film. ͑ р ͒ ϭ pinned while H HB on linear defects, and Jc(H) 5 —The ‘‘motional narrowing’’ of the damping peak is ϫ109 –1010 A/m2. Therefore, when some vortices are found indicative of ordering of the vortex lattice upon the passage inside the domains and become unstable to lateral deforma- of current. The remarkable agreement between the measured tions owing to the sharp decrease of their effective linear damping peak and the theoretical curve obtained in the tension, Jc(H) does not increase but instead begins to fall. framework of the diffusion model attests to preservation of a Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan 745 gradual transition ͑the absence of sharp phase transforma- found gratitude to their colleagues and friends E. A. tions͒ upon the depinning of the vortices in systems with Pashitskii, A. L. Kasatkin, V. F. Solov’ev, V. L. Svechnikov, strong disorder. V. S. Flis, V. A. Komashko, H. W. Zandbergen, G. W. Crab- —The ‘‘vortex glass’’transition temperature obtained by tree, P. Esquinazi, C. L. Snead, M. Suenaga, and M. Lorenz, means of scaling of the resistivity coincides with the depin- who all contributed immeasurably to the development of this ning temperature measured at zero current, indicating that paper. The authors are grateful to T. H. Geballe, J. R. Clem, the interpretation of this transition is ambiguous and might E. H. Brandt, D. Dew-Hughes, D. Larbalestier, H. Ku¨pfer, be explained in terms of a model of vortex depinning. and A. A. Zhukov for helpful discussions. —Linear defects oriented along the c axis, i.e., disloca- tion ensembles a ͓100͔, are the reason for the appearance of *E-mail: [email protected] ʈ the Jc(H c) peak, while dislocations and loops lying in the **E-mail: [email protected] ʈ ab plane contribute to the peak Jc(H ab) and also to the ʈ peak Jc(H c) as a result of the interaction of vortices with transverse dislocation lines as with pointlike defects. 1 —Opposite temperature behavior of the peaks of L. V. Shubnikov, V. I. Khotkevich, Yu. D. Shepelev, and Yu. N. Ryabinin, Zh. E´ ksp. Teor. Fiz. 7, 221 ͑1937͒. ␪ ʈ ʈ 2 Jc(H, ) for H c and H ab is found experimentally. The A. A. Abrikosov, Zh. E´ ksp. Teor. Fiz. 32,1442͑1957͓͒Sov. Phys. JETP 5, ʈ ͑ ͔͒ peak Jc(H c) initially grows with increasing temperature 1174 1957 . 3 V. V. Shmidt, Introduction to the Physics of Superconductors ͓in Russian͔, and then is suppressed when T approaches Tirr(H). In con- ͑ ͒ ʈ Nauka, Moscow 1982 ,p.240. trast, the relative height of the peak Jc(H ab) becomes 4 U. Essman and H. Tra¨uble, Phys. Lett. A 24,526͑1967͒. larger and larger as the temperature increases. This behavior 5 V. Vinokur, B. Khaykovich, E. Zeldov, M. Konczykowski, R. A. Doyle, is due to the different temperature dependence of the differ- and P. H. Kes, Physica C 295, 209 ͑1998͒. 6 ent contributions to J (Hʈc) and J (Hʈab): pinning on dis- G. Blatter, M. V. Feigel’man, V. B. Geshkenbein, A. I. Larkin, and V. M. c c Vinokur, Rev. Mod. Phys. 66, 1125 ͑1994͒. locations, pinning on pointlike defects, and the effect of the 7 M. P. A. Fisher, Phys. Rev. Lett. 62, 1415 ͑1989͒. anisotropy of the strain fields near the dislocation cores. 8 D. S. Fisher, M. P. A. Fisher, and D. A. Huse, Phys. Rev. B 43,130͑1991͒. 9 —In the neighborhood of the cores of edge dislocations A. E. Koshelev and V. M. Vinokur, Preprint cond-mat/9801144 ͑January 14, 1998͒. whose lines are parallel to the c axis a strain field arises 10 G. W. Crabtree and D. R. Nelson, Phys. Today 50,38͑1997͒. which plays an important role in the formation of the pinning 11 D. R. Nelson and V. M. Vinokur, Phys. Rev. B 48, 13060 ͑1993͒. potential and the local suppression of the superconducting 12 V. F. Solovjov, V. M. Pan, and H. C. Freyhardt, Phys. Rev. B 50, 13724 ͑ ͒ order parameter and T . For TϽT this will give rise to 1994 . c c 13 V. M. Pan, V. F. Solovjov, A. L. Kasatkin et al.,inPhysics and Materials anisotropic regions of the normal phase near the dislocation Science of High Temperature Superconductivity IV, Vol. 26 of NATO ASI core. Series, edited by R. Kossowsky et al., Kluwer Academic Publ., Dordrecht, —In the neighborhood of dislocation lines parallel to the Boston, London ͑1997͒,p.3. 14 V. M. Pan, in Physics and Materials Science of Vortex States, Flux Pin- ab plane no appreciable strain fields appear. It can be as- ning and Dynamics, Vol. 356 of NATO ASI Series, edited by R. sumed that this is the cause of the opposite behavior of the Kossowsky et al., Kluwer Academic Publ., Dordrecht, Boston, London pinning force and critical current density in the case ͑1999͒,p.1. 15 J (Hʈab). V. M. Pan, V. F. Solovjov, and H. C. Freyhardt, in Advances in Cryogenic c Engineering, Materials, Vol. 42, edited by L. T. Summers, Plenum Press, ʈ —The measured Jc(H) curves for H c in YBCO films New York ͑1997͒, p. 663. demonstrate two clear crossovers, corresponding to transi- 16 V. M. Pan, V. F. Solovjov, and H. C. Freyhardt, Physica C 279,18͑1997͒. tions from J (H)ϳconst at low fields, in which the vortex 17 V. M. Pan, V. F. Solovjov, and H. C. Freyhardt, Czech. J. Phys. 46,1643 c ͑ ͒ lattice has not yet formed and the interaction of vortices with 1996 . 18 V. M. Pan, Usp. Fiz. Met. 1,49͑2000͒. linear centers occurs in the regime of individual pinning, to 19 T. Hwa, D. R. Nelson, and V. M. Vinokur, Phys. Rev. B 48,1167͑1993͒. ϰ Ϫ0.5 ͑ 20 Jc H at intermediate fields the motion of the vortex J. Mannhart, D. Anselmetti, J. G. Bednorz et al., Semicond. Sci. Technol. ͒ 5,125͑1992͒. lattice as a deformable 2D continuum and, finally, to Jc 21 ϰ Ϫq ͑ ϳ ͒ D. G. Schlom, D. Anselmetti, J. G. Bednorz et al., Z. Phys. B: Condens. H where q 1 – 1.5 at higher fields. Matter 86, 163 ͑1992͒. —The inhomogeneous state of a vortex ensemble inter- 22 V. L Svetchnikov, Met. Phys. Adv. Tech. ͑2001͒, in press. acting with an ensemble of parallel equidistant linear centers 23 S. K. Streiffer, B. M. Lairson, C. B. Eom et al.,Phys.Rev.B43, 13007 ͑ ͒ which are arranged in a network substantially differs from 1991 . 24 S. J. Pennycook, M. F. Chisholm, D. E. Jensson et al., Physica C 202,1 the state of a random 2D Bose glass. ͑1992͒. This study was done with the support of the State fund 25 V. Svetchnikov, V. Pan, Ch. Traeholt, and H. Zandbergen, IEEE Trans. for Basic Research of the Ministry of Education and Science Appl. Supercond. AS-7, 1396 ͑1997͒. ͑ ͒ 26 V. Narlikar and D. Dew-Hughes, Phys. Status Solidi 6, 383 ͑1964͒. of Ukraine, project No. 2–4/349 contract No. F4/147–97 , 27 M. F. Chisholm and D. A. Smith, Philos. Mag. A 59, 181 ͑1989͒. the International Program for Support of Science and Educa- 28 Y. Gao, K. L. Merkle, G. Bai, H. L. M. Chang, and D. J. Lam, Physica C tion ͑ISSEP͒, grants SPU 062044 and SPU 072046, the in- 174,1͑1991͒. 29 ternational association INTAS, grant No. 99-00585, the Fed- K. Merkle, Interface Sci. 2,311͑1995͒. 30 V. Hardy, J. Provost, D. Groult et al., J. Alloys Compd. 195,395͑1993͒. eral Ministry of Education, Science, Research, and 31 S.-W. Chan, J. Phys. Chem. Solids 55,1415͑1994͒. Technology of Germany ͑BMBF FKZ TRANSFORM͒, grant 32 D. R. Nelson and V. M. Vinokur, Phys. Rev. Lett. 68,2398͑1992͒. 13 N 7218/7 through the Institute of Experimental Physics II 33 A. Gervais and D. Keller, Physica C 246,29͑1995͒. 34 of the University of Leipzig, and also the Department of Z. L. Wang, D. H. Lowndes, D. H. Christen, D. K. Kroeger, C. E. Klabunde, and D. P. Norton, Physica C 252, 125 ͑1995͒. Energy of the USA through Brookhaven National Labora- 35 V. M. Pan, A. L. Kasatkin, V. L. Svetchnikov, and H. W. Zandbergen, tory, contract No. 851924. The authors express their pro- Cryogenics 33,21͑1993͒. 746 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Pan and A. V. Pan

36 V. M. Pan, A. L. Kasatkin, and H. C. Freyhardt, IEEE Trans. Appl. Su- 65 K. Yamafuji and T. Kiss, Physica C 258,197͑1996͒. percond. AS-7, 1588 ͑1997͒. 66 T. Matsushita, T. Tohdoh, and N. Ihara, Physica C 259, 321 ͑1991͒. 37 A. L. Kasatkin and V. M. Pan, The Ninth International Workshop on 67 M. Ziese, Phys. Rev. B 53, 12422 ͑1996͒; 55, 8106 ͑1997͒; Physica C 269, Critical Currents, IWCC 9-99, Extended Program Book, July 7–10, 1999, 35 ͑1996͒. The Pyle Center, University of Wisconsin-Madison, USA ͑1999͒,p.14. 68 B. Roas, L. Schultz, and G. Saemann-Ischenko, Phys. Rev. Lett. 64,479 38 E. H. Brandt, Phys. Rev. Lett. 69, 1105 ͑1992͒. ͑1990͒. 39 V. M. Pan, V. A. Komashko, V. S. Flis, A. L. Kasatkin, V. L. Svetchnikov, 69 M. Fendorf, C. P. Burmester, L. T. Wille, and R. Gronsky, Appl. Phys. O. P. Karasevskaja, A. G. Popov, A. V. Pronin, C. L. Snead, M. Suenaga, Lett. 57, 2481 ͑1990͒. and H. W. Zandbergen, Physica B 284–288, 831 ͑1999͒. 70 V. Svetchnikov, A. Palti, and V. Pan, Met. Phys. Adv. Tech. 17,257 40 J. Mannhart, J. Supercond. 3, 281 ͑1990͒. ͑1998͒. 41 R. Gross and B. Mayer, Physica C 180, 235 ͑1991͒. 71 V. Svetchnikov, V. Vysotskii, and V. Pan, Met. Phys. Adv. Tech. 17, 1235 42 P. A. Nilsson, D. Winkler, J. A. Alarco et al., Appl. Phys. Lett. 59, 3030 ͑1999͒. ͑1991͒. 72 V. M. Pan, A. L. Kasatkin, V. L. Svetchnikov, V. A. Komashko, A. G. 43 D. Shi, Appl. Supercond. 1,61͑1993͒. Popov, A. Yu. Galkin, H. C. Freyhardt, and H. W. Zandbergen, IEEE 44 D. Dimos, P. Chaudhari, and J. Mannhart, Phys. Rev. B 41, 4038 ͑1990͒. Trans. Appl. Supercond. AS-9, 1532 ͑1999͒. 45 A. Gurevich, H. Kuepfer, and C. Keller, Semicond. Sci. Technol. 4,91 73 V. Selvamanickam, M. Mironova, S. Son, and K. Salama, Physica C 208, ͑1991͒. 238 ͑1993͒. 46 A. Gurevich, Phys. Rev. B 42,4857͑1990͒. 74 V. M. Pan, V. S. Flis, O. P. Karasevska, V. I. Matsui, I. I. Peshko, V. L. 47 M. Strikovsky, G. Linker, S. V. Gaponov et al., Phys. Rev. B 45, 12522 Svetchnikov, M. Lorenz, A. N. Ivanyuta, G. A. Melkov, E. A. Pashitskii, ͑1992͒. and H. W. Zandbergen, J. Superconductivity and Incorporated Magnetism 48 A. V. Pan, F. Ciovacco, P. Esquinazi, and M. Lorenz, Phys. Rev. B 60, 14,109͑2001͒. 4293 ͑1999͒. 75 V. M. Pan, V. S. Flis, V. A. Komashko, O. P. Karasevska, V. L. Svetchni- 49 P. Esquinazi, J. Low Temp. Phys. 85, 139 ͑1991͒. kov, M. Lorenz, A. N. Ivanyuta, G. A. Melkov, E. A. Pashitskii, and H. W. 50 M. Ziese, P. Esquinazi, and H. F. Braun, Semicond. Sci. Technol. 7, 869 Zandbergen, IEEE Trans. Appl. Supercond. AS-11,3960͑2001͒. ͑1994͒. 76 V. M. Pan, A. L. Kasatkin, V. S. Flis, V. A. Komashko, V. L. 51 E. H. Brandt, Phys. Rev. Lett. 68, 3769 ͑1992͒. Svetchnikov, A. G. Popov, A. V Pronin, O. P. Karasevska, C. L. Snead, M. 52 M. Ziese, P. Esquinazi, Y. Kopelevich, and A. B. Sherman, Physica C 224, Suenaga, and H. W. Zandbergen, Institute of Physics Conference Series 79 ͑1994͒. No. 167, Applied Superconductivity ͑2000͒,p.699. 53 F. Pardo, F. de la Cruz, P. Gammel, E. Bucher, and D. J. Bishop, Nature 77 V. A. Komashko, A. G. Popov, V. L. Svetchnikov, A. V. Pronin, V. S. ͑London͒ 396,348͑1998͒. Melnikov, A. Yu. Galkin, V. M. Pan, C. L. Snead, and M. Suenaga, Semi- 54 P. Le Doussal and T. Giamarchi, Phys. Rev. B 57, 11356 ͑1998͒. cond. Sci. Technol. 13, 209 ͑2000͒. 55 S. Scheidel and V. Vinokur, Phys. Rev. B 57, 2574 ͑1998͒. 78 H. Safar, J. Y. Coulter, M. P. Maley et al., Phys. Rev. B 52,9875͑1995͒. 56 L. N. Bulaevskii, M. Ledvij, and V. G. Kogan, Phys. Rev. B 46, 366 79 B. Meingast, O. Krant, T. Wolf, H. Wuehl, A. Erb, and G. Mueller-Vogt, ͑1992͒. Phys. Rev. Lett. 67, 1639 ͑1991͒. 57 A. V. Pan, R. Hohne, M. Ziese, P. Esquinazi, and C. Assmann, in Physics 80 U. Welp, M. Grimsditch, S. Fleshler, W. Nessler, B. Veal, and G. W. and Materials Science of Vortex States, Flux Pinning and Dynamics, Vol. Crabtree, J. Supercond. 7, 159 ͑1994͒. 356 of NATO Science Series, edited by R. Kossowsky et al., Kluwer 81 B. Meingast, A. Junod, and E. Walker, Physica C 272, 106 ͑1996͒. Academic Publ., Dordrecht, Boston, London ͑1999͒, p. 545. 82 A. Gurevich and E. A. Pashitskii, Phys. Rev. B 56, 6213 ͑1997͒. 58 A. V. Pan, PhD thesis, University of Leipzig, Leipzig ͑2000͒. 83 J. B. Hirth and J. Lothe, Theory of Dislocations, McGraw-Hill, New York 59 V. M. Pan, G. G. Kaminsky, A. L. Kasatkin, M. A. Kuznetsov et al., ͑1988͒. Supercond. Sci. Technol. 5,48͑1992͒. 84 H. Ledbetter and M. Lei, J. Mater. Res. 6, 2253 ͑1991͒. 60 H. Jensen et al., Phys. Rev. Lett. 60, 1676 ͑1988͒; Phys. Rev. B 38, 9235 85 B. Dam, J. M. Huijbregtse, F. C. Klaassen, R. C. F. van der Geest, G. ͑1988͒. Doornbos, J. H. Rector, A. M. Testa, S. Freisem, J. C. Martinez, B. 61 A. V. Pan, P. Esquinazi, and M. Lorenz, Phys. Status Solidi B 215,573 Staeuble-Puempin, and R. Griessen, Nature ͑London͒ 399,439͑1999͒. ͑1999͒. 86 J. M. Huijbregtse, F. C. Klaassen, R. C. F. van der Geest, B. Dam, and 62 E. H. Brandt, Rep. Prog. Phys. 58, 1465 ͑1995͒. R. Griessen, J. Low Temp. Phys. 32,114͑1999͒. 63 R. H. Koch, V. Foglietti, W. J. Gallagher, G. Koren, A. Gupta, and M. P. A. 87 L. Civale, A. D. Marwick, T. K. Worthington, M. A. Kirk, J. R. Thompson, Fisher, Phys. Rev. Lett. 63,1511͑1989͒; 64, 2586 ͑1990͒. L. Krusin-Elbaum, Y. Sun, J. R. Clem, and F. Holtzberg, Phys. Rev. Lett. 64 J. R. Thompson, Y. R. Sun, L. Civale, A. P. Malozemoff, M. W. 65,648͑1991͒. McElfresh, A. D. Marwick, and F. Holtzberg, Phys. Rev. B 47, 14440 ͑1993͒. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Magnetic properties of irradiated quasi-2D type-II superconductors G. M. Braverman

Max Planck Institut fu¨r Kernphysik, Heidelberg, Germany S. A. Gredeskul* and Y. Avishai

Department of Physics, Ben-Gurion University of the Negev, Beer-Sheva 84105, Israel ͑Submitted April 3, 2001͒ Fiz. Nizk. Temp. 27, 1012–1018 ͑September–October 2001͒

Persistent scaling behavior of magnetization in layered high Tc superconductors with short-range columnar defects is explained within the Ginzburg–Landau theory. In the weak field region, the scaling function differs from that of a clean sample and the critical temperature is renormalized due to defects. In the strong field region, defects are effectively suppressed and the scaling function and critical temperature are the same as in a clean superconductor. This picture is consistent with recent experimental results. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1401183͔

1. INTRODUCTION Moreover, putting this factor equal to a constant from the Layered high-temperature superconducting ͑HTSC͒ ma- very beginning, Tesanovic´ et al. obtained an explicit form of 9 terials, such as Bi2Sr2CaCu2O8ϩ␦ and Bi2Sr2Ca2Cu3O10, are the scaling function, known experimentally to exhibit 2D scaling of magnetic 1 1,2 2 properties around the mean-field transition line Hc2(T). It ␥ ͑x͒ϭ ͑ͱx ϩ2Ϫx͒ ͑2͒ 0 2 is manifested by inspecting the magnetization M 0 as a func- ͑ ͒ tion of temperature T and the external magnetic field H: which agrees, to good accuracy, with the experimental mag- s⌽ netization data for Bi2Sr2Ca2Cu3O10. Such a form of the 0 ͑ ͒ϭϪ ␥ ͑ ͒ ͑ ͒ ͑ ͒ ͱ M 0 T,H 2 0 x , 1 scaling function 2 implies the existence of a crossing point: A kBTH ϭ ϩ 2 Ј Ϫ1 at some temperature T0* Tc0(1 kB /(2A Hc2)) the where s is an effective interlayer spacing, ⌽ is the flux 0 sample magnetization is independent of H, M 0* quantum, xϭAHЈ ͓TϪT (H)͔/ͱk TH is the scaling vari- ϵ ϭϪ ⌽ c2 c2 B M 0(T0* ,H) kBT0*/(s 0). Later on, Tesanovic´ and ␥ Ј 10 ␤ able, 2 0(x) is so called scaling function, Hc2 Andreev took the fluctuations of A into account and gen- ϵϪdH (T)/dT͉ ϭ , and T is the zero-field critical c2 T Tc0 c0 eralized the approach developed in Ref. 9 to arbitrary type-II temperature. For a superconductor with Ginzburg–Landau superconductors. ͑ ͒ ␬ 3 ␤ GL parameter and Abrikosov geometric factor A the Recently the influence of columnar defects on the mag- ϭͱ ⌽ ϭ ␲␬2␤ constant A s 0 /p, where p 16 A . netic properties of superconductors has been studied ␥ 11,12 13–17 The scaling function 0(x) was first evaluated in the experimentally and theoretically. Columnar defects perturbative regime4,5 for xӶ1. A nonperturbative result for emerge after heavy ion irradiation of the superconducting it was obtained later using the following arguments. The sample.11 They serve as strong pinning centers: each one is compounds mentioned above are strongly type-II supercon- able to pin a single vortex. The radius of a columnar defect ductors with large GL parameters ␬Ϸ100. Their effective can be larger or smaller than the coherence length ͑long- interlayer separation sϭ1.5 nm is larger than the effective range or short-range defects, respectively͒. Strong columnar superconducting coherence length ␰(H,T) ͑if the magnetic defects lead to the formation of multiquantum vortices in field is not extremely close to the mean field transition field high-temperature superconductors13,14 and in conventional ͒ 16 Hc2(T) , but is much smaller than the magnetic penetration ones as well. They also lead to additional magnetization 17 depth. Hence the problem of fluctuations near Hc2(T) be- jumps in mesoscopic samples. Therefore it is interesting to comes effectively two dimensional and can be represented understand how such defects influence both the scaling be- theoretically in terms of the 2D GL mean-field theory pro- havior and the existence of the crossing point. jected onto the lowest Landau level ͑LLL͒.6 Such an ap- Specifically we refer to experiments performed by van Ͼ 12 proximation remains valid at least for H Hc2(T)/3, when der Beek et al., who studied the thermodynamic properties higher Landau levels are obviously irrelevant. Moreover, re- of single crystals of Bi2Sr2CaCu2O8ϩ␦ . The samples were cent results7 show that LLL projection is valid even for irradiated with 5.8-GeV ions that produced columnar defects Ͼ 6 ϭ ϭ ϫ 10 Ϫ2 H Hc2(T)/13. Tesanovic´ emphasized the crucial role with radius L 3.5 nm and 2D density nd 5 10 cm . played by the total amplitude of the order parameter in the This density is small in the sense that the matching field ϭ ⌽ critical region. Integration of the partition function over this H⌽ nd 0 , at which the number of vortices becomes equal ␤ amplitude, assuming that the Abrikosov factor A depends to the number of defects, is much smaller than Hc2(T). The weakly on the vortex configuration, leads8 to the scaling law magnetization was measured in the region of magnetic fields ͑ ͒ ͑1͒. 0.2–5 T which are also smaller than Hc2(T) and tempera-

1063-777X/2001/27(9–10)/5/$20.00747 © 2001 American Institute of Physics 748 Low Temp. Phys. 27 (99–10), September–October 2001 Braverman et al. tures 72–86 K. Within this interval of fields the defect radius Abrikosov factor almost coincides with its ‘‘triangular’’ ␤ ϭ is the smallest length scale in the problem, and the defects value A 1.16. can be treated as short-range. Measurements showed that co- The next Section contains the main body of the paper. lumnar defects drastically change the reversible magnetiza- First we formulate the model ͑subsection 2.1͒ and then ob- tion of the sample: there are now two scaling regimes, per- tain the magnetization in the weak field region ͑subsection taining to relatively weak (HϽH⌽) and strong (HϾH⌽) 2.2͒ and in the strong field region ͑subsection 2.3͒. Relation magnetic fields. These two regimes are described by the of our calculations to the experiment12 is discussed in Sec. 3. same form ͑2͒ of the scaling function as for clean sample but Finally, the results obtained in this work are summarized in they correspond to two different zero-field critical tempera- Sec. 4. tures ͑used in Ref. 12 as fitting parameters͒ and two crossing points. 2. SCALING BEHAVIOR OF THE IRRADIATED In this paper we propose an explanation of these results. SUPERCONDUCTOR Our arguments are based on the observation that the mag- 2.1. The model netic field serves as a control parameter for tuning an effec- ͑ tive concentration cϭH⌽ /H of defects ͑the number of de- Consider an irradiated thin superconducting film or one fects divided by the number of vortices͒. In the weak field layer in a layered superconductor͒ with area S subject to a ͑ ϭ ␮ ϭ perpendicular magnetic field ͑thus parallel to the defects͒. region the concentration is large c 5 for 0H 0.2 T and n from Ref. 12͒, and each vortex is affected by a force Columnar defects can be described as a local reduction of the d ␦ ϭ ⌺ Ϫ Ϫ 2 2 emanating from many defects. On the average, this force critical temperature Tc(r) Tc0 t j exp( (r rj) /2L ). Here r is a two-dimensional vector in the film plane, L is the leads to renormalization of the critical temperature Tc . Short-range defects are effectively weak and can be taken defect radius, and the positions rj of defects are uniformly into account perturbatively. In first order they retain the same and independently distributed over the film plane with den- form of the scaling function ͑2͒ as that of a clean sample, up sity nd . The value of nd is assumed to be moderate, so that to the aforementioned renormalization of the critical tem- for the pertinent region of temperature the matching field H⌽ perature. Second-order corrections indeed destroy the scaling is always much smaller than Hc2(T). The dimensionless am- р р behavior, but in the vicinity of the crossing temperature scal- plitudes of the defects 0 t j 1 are also independent random ing is approximately maintained. In the strong field region quantities distributed with some probability density p(t) whose first two moments ␪ and ␪ satisfy 0р␪ р1. On the concentration is small ͑cϭ0.2 for ␮ Hϭ5 T and n from 1 2 1,2 0 d average, the defects lead to renormalization of the critical Ref. 12͒, renormalization is not needed, and the standard temperature concentration expansion18 can be used. Here, strictly speak- ͑ ϭ Ϫ␦ ͑ ͒ ing, even the first-order correction with respect to small con- Tc Tc0 Tc , 3 centration͒ destroys the scaling behavior. However, a strong where field effectively suppresses the defects, thus restoring the ␦ ϵ ␦ ͒ ϭ ␲ 2␪ scaling behavior of a clean superconductor with the initial Tc ͗ Tc͑r ͘ 2 ndL 1Tc0 . critical temperature Tc0 . Identifying the two fitting tempera- The fluctuation of the shift of the critical temperature has tures of Ref. 12 with the renormalized and initial critical zero mean value and a variance temperatures Tc and Tc0 , respectively, one finds for the di- ␪ ϭ ͗͑␦T ͑r͒Ϫ␦T ͒2͘ϭ␲n L2␪ T2 . mensionless defect strength 1 0.49, well within its allowed c c d 2 c0 р␪ р range 0 1 1. This indicates complete consistency be- The thermodynamic properties of a type-II supercon- tween the description constructed below and the experimen- ductor are described by its partition function tal results of Ref. 12. Our quantitative approach follows the one proposed and G Zϰ ͵ D͕⌿͖expͩ Ϫ ͪ , ͑4͒ successfully used within the critical region in clean kBT 8–10 superconductors and at low fields in disordered ⌿ 15 where is the order parameter and G is the standard GL superconductors. This approach is based on the LLL pro- functional jection and on the assumption that the Abrikosov factor is ␤ ⌿ͯ2 ץͯ␥almost independent of the magnetic field. The latter assump- ϭ ͵ ͭ ␣͑ ͒ͯ⌿ͯ2ϩ ͯ⌿ͯ4ϩ G s r Ϫ tion is evidently not valid in the vicinity of the matching 2 field, but it is valid for fields much smaller or much larger 1 than H⌽ . Indeed, columnar defects are strong pinning cen- ϩ ͑BϪH͒2ͮ d2r, ͑5͒ ters. Then, if the number of vortices is much less than or 8␲ much greater than the number of defects, configurations ប 2e ,Ϫϭ ٌϩ A, BϭٌϫAץ -close to the triangular Abrikosov lattice are always simulta neously compatible with any typical configuration of defects i c and with the condition of complete ͑as possible͒ pinning. with the first GL coefficient ␣(r) depending on coordinate This assumption is supported by noticing the remarkable dif- through a local change of the critical temperature. ference between the number of vortices and the number of Further simplifications will be done for the case of weak defects in both regions of fields. This enables us to take into fields ͑in the case of strong fields slightly different simplifi- account only these Abrikosov-like vortex configurations and cations are required—see subsection 2.3 below͒. Let us take fluctuations around them.1͒ But for these configurations the into account the large value of the GL parameter, project the Low Temp. Phys. 27 (99–10), September–October 2001 Braverman et al. 749

→ϱ system on the LLL corresponding to the external field, intro- In the thermodynamic limit S with nd and Nv /S duce the scaling variable x as mentioned above, a scaled fixed, the partition function ͑9͒ could be calculated in a order parameter ␸ϰ⌿, and dimensionless temperature fluc- saddle-point approximation. This results in the following tuations form for the magnetization:

⌽ 1/2 ͒ ␦ ͒Ϫ␦ ⌽ s 0 Hc2͑T Tc͑r Tc s 0 ␶͑r͒ϭͩ ͪ . M͑T,H͒ϭϪ2␥͑x͒, ͑12͒ p ͱ T ϪT ͱ kBTH c A HT ␥ ͑We emphasize that for the case of weak fields it is the where (x) is the solution of the saddle-point equation ϭ0. ͑13͒␥ץ/L͑␥,x͒ץ -renormalized critical temperature which enters the expres sion for the upper critical field as well as the definition of the ␶ϭ scaling variable x.͒ This results in the following expression In the case ˆ 0 the two possible saddle points satisfy for the partition function: the equation 1 1 Ϫ ␥ϩ ϭ Zϰ ͵ D͕␸͖ ͭ Ϫ ͩ ͉␸͉2ϩ ͉␸͉4ϩ␶͑ ͉͒␸͉2 ͪͮ 2 ϩ␥ 0, exp Nv x ␤ r , x 4 A ͑6͒ but only one of them ͑ where Nv is the total number of vortices i.e., the total num- 1 ͒ ␥ ͑x͒ϭ ͑ͱx2ϩ2Ϫx͒ ͑14͒ ber of flux quanta through the sample area S and the bar 0 2 denotes averaging over the sample area. The expression for ⌫ the magnetization has the form can be reached by an allowed deformation of the contour . Substitution of Eq. ͑14͒ into ͑12͒ yields the magnetization ͑ Z ץ ⌽ s 0 1 ln M 0(T,H) of a clean sample up to renormalization of the M͑T,H͒ϭ . ͑7͒ ͑ ͒ ͑ ͒͒ x critical temperature 1 , 2 obtained in Ref. 9. The saddleץ ͱ N A kBTH v ␥ point 0(x) serves as the scaling function. These two formulas ͑6͒, ͑7͒ form the basis for the further Returning to the disordered case, we note that in the calculation and analysis. thermodynamic limit the last term on the right-hand side of ͑ ͒ Ϫ1 Eq. 10 has an explicit self-averaged structure Nv tr(...) and can therefore be replaced by its average. This procedure modifies the saddle-point equation to 2.2. Magnetization: weak fields 1 ␧͑T͒ The assumption that the Abrikosov factor is a constant ␥ϭ ϩ ͑ ͒ 2 ϩ␥ ͑ ϩ␥͒3 15 enables us, following Ref. 15, to replace ͉␸(r)͉4 in Eq. ͑6͒ x x ␤ ͉␸ ͉2 2 by A( (r) ) . This replacement, together with the sim- and results in a magnetization plest version of the Hubbard-Stratonovich transformation ͑ 2␥ ͑x͒ introduction of an additional integration over some auxiliary ͑ ͒ϭ ͑ ͒ͩ ϩ␧͑ ͒ 0 ͪ ͑ ͒ M T,H M 0 T,H 1 T , 16 field ␥͒ turns the problem to be an exactly solvable one.15 ͱx2ϩ2 Indeed, we expand the order parameter on the LLL subspace, where Nv ␶2 ␪ ͑ ␲ Ј ͒2 2 ␸͑ ͒ϭ ͑ ͒ ͑ ͒ trˆ 2 2 Hc2Tc0 sL r ͚ CmLm r , 8 ␧͑ ͒ϭͳ ʹ ϭ 2 ͑ ͒ mϭ0 T ndL . 17 Nv p kBT where Lm(r) are normalized LLL eigenfunctions with orbital Note that the parameter ␧(T) is proportional to the fourth momentum m. Then after integration over the expansion co- power of the defect radius L, thus justifying the perturbation ͑ ͒ efficients Cm , the partition function 6 reads approach for short-range defects. In the zeroth-order approximation with respect to ␧(T) Zϰ ͵ ͕Ϫ L͑␥ ͖͒ ␥ ͑ ͒ exp NV ,x d , 9 the magnetization has exactly the same form as for a clean ⌫ sample, thus retaining both the scaling property and the ex- where istence of a crossing point. However, due to renormalization of the critical temperature, the crossing temperature T* L͑␥ ͒ϭϪ␥2ϩ Ϫ1 ͓͑ ϩ␥͒ˆϩ␶͔ ͑ ͒ ϭ *Ϫ␦ * ,x Nv tr ln x l ˆ 10 T0 T* differs from its value T0 for a clean sample without defects: ␦T*ϭ␦T (1ϩ(2A2HЈ )Ϫ1)Ϫ1. In the next ␶ c c2 and ˆ is a random matrix with elements order, scaling is virtually destroyed since the correction term ͑within the parenthesis in Eq. ͑16͒͒ depends not only on the ␶ ϭ ͵ ͑ ͒␶͑ ͒ ͑ ͒ 2 ͑ ͒ mn Lm* r r Ln r d r. 11 scaling variable x but also on temperature. But at the cross- S ing temperature T* the magnetization reads ⌫ ͑ ͒ The contour in Eq. 9 is parallel to the imaginary axis and 2H* ␥ Ϫ ϱ ␥ ϩ ϱ ͑ ͒ϭ ͑ ͒ͩ ϩ␧͑ ͒ ͪ ͑ ͒ stretches from * i to * i . To assure convergence of M T*,H M 0 T* 1 T* ϩ , 18 ␥ H H* the integrals over the coefficients ͕Cm͖ the real constant * ␥ ϩ ϩ ␶ Ͼ ␶ ϭ ϭ 2 should satisfy the inequality * x min n 0, where n is where H* Hc2(T*) kBT*/(2A ). Therefore if the field is ␶ Ӷ the nth eigenvalue of the matrix mn . weak enough, H H*, then the crossing point is restored, 750 Low Temp. Phys. 27 (99–10), September–October 2001 Braverman et al.

T* serves as a true crossing temperature, and the magnetiza- strong field region H⌽ӶHӶH* the crossing temperature co- tion at the crossing temperature differs from its unperturbed incides with its initial value T0* and the magnetization in the Ϫ ␥ form 2 0(x) merely by a multiplicative constant crossing point practically coincides with its value M 0* in a 1ϩ2␧(T*). clean superconductor. 2.3. Magnetization: strong fields 3. DISCUSSION When the magnetic field increases, the approach used above becomes inapplicable. First, it fails in the vicinity of According to the results obtained above, within the main the matching field, where the Abrikosov factor becomes very approximation, the magnetization indeed manifests two sepa- sensitive to the details of the defect configuration. Secondly, rate scaling regimes in the regions of weak and strong mag- higher-order terms in the perturbation expansion for the netic fields. These regimes are described by the same scaling saddle-point equation ͓which were omitted in Eq. ͑16͔͒ grow function that characterizes a clean sample, but with a renor- with magnetic field. Fortunately, we have here a new small malized critical temperature in the weak region and with the parameter, because the dimensionless concentration c of de- initial critical temperature at high fields. These results are in fects in the strong field region is small. Therefore there is no complete qualitative correspondence with the experimental sense in renormalizing the critical temperature, and it is natu- observations of van der Beek et al. Without pretending to ral to use the concentration expansion.18 Then in this region account for a complete quantitative description of the above the dimensionless temperature fluctuation ␶(r) is now de- mentioned experiments, we nevertheless will show that some fined as quantitative agreement can also be achieved. Let us discuss the limits of applicability of our results s⌽ 1/2 H ͑T͒ ␦T ͑r͒ and their relation to the experiment of Ref. 12. ␶͑ ͒ϭ 0 c2 c ͑ ͒ r ͩ ͪ Ϫ . 19 1. In the weak field region, the important small param- p ͱk TH Tc0 T B eters are ␧(T) ͑which enters the magnetization ͑18͒͒ and ␧ ϩ␥ 2 ͑ ͒ As mentioned above, the second term in the right-hand (T)/(x 0(x)) which enters the saddle-point equation . side of Eq. ͑10͒ is self-averaging and can be calculated using Using the parameters which were employed in the experi- 12 ͑ ␮ Ј ϭ Ϫ1 ϭ the limiting form of the density of states ␳͑␶͒ of the matrix mental work kB 0Hc2 1.15 T•K , T* 78.9 K, and the ͒ ͑11͒. For short-range defects in the linear approximation with rest of parameters which were already mentioned in Sec. 1 , ͑ ͒␧ ϭ ␪ ␧ respect to c, this density of states reads we find from Eq. 17 (T*) 0.5 2 and (T*)/(x* ϩ␥ 2Ϸ ͑ ␮ 0(x*)) 0.25 the latter figure is obtained for 0H ϭ ͒ ␪ ϭ c ␶ 0.2 T . For the quite plausible value 2 0.5 one then finds ␳͑␶͒ϭ͑1Ϫc͒␦͑␶͒ϩ pͩ ͪ , ͑20͒ ␧ ϭ ␪ ϭ ␭ ␭ (T*) 0.5 2 0.25. 2. The condition of convergence of the integral over the ␭ϭ ␲ 2 Ј ͱ ⌽ ͱ Ϫ1 ͕ ͖ where 2 L Tc0AHc2 H( 0 kBT) and p(t) is the expansion coefficients Cm can be written as H Ͼ ␪2 ␪ ␪2ϭ␪ probability distribution of the dimensionless temperature t j . 0.25H⌽ 1/ 2 , and even in the worst case 1 2 it reads ␶ ␮ Ϸ Indeed, the matrix mn is nothing but the Hamiltonian of a 0H 0.25 T. ␮ Ϸ particle with charge 2e in a 2D system subject to a perpen- 3. Then, one has 0H* 6.4 T, and applicability of the ␮ Ͼ dicular magnetic field and containing short-range defects LLL projection requires 0H 0.5 T. The weak field region ␮ ϭ ͑projected on the LLL͒. The first and second terms in Eq. of Ref. 12 corresponds to 0H 0.2-0.02 T. Thus, in the ͑20͒ correspond, respectively, to those states whose energy is weak field region, only the condition for applicability of the stuck to the LLL ͑despite the presence of zero-range defects LLL projection is violated slightly, but the deviation is not ͑see, e.g., Ref. 19͒͒ and those states whose energies are lifted dramatic. from the LLL by these defects. For sufficiently narrow dis- 4. In the strong field region we find ␩Ϸ2.9, and there- tribution p(t), the corresponding saddle-point equation leads fore the correction term in parenthesis in Eq. ͑22͒ is less than to the magnetization three percent, so that in this region our assumptions are fully satisfied. c␭␪ Hence, up to an insignificant mismatch for very weak ϭ ͫ Ϫ 1 ͬ ͑ ͒ M M 0 , 21 fields our theoretical assumptions and simplifications are ͑1ϩ2␭␪ ␥ ͑x͒͒ͱx2ϩ2 1 0 completely consistent with the experimental parameters of ͑ ͒ Ref. 12. Using the same set of parameters, we display in Fig. where M 0(T,H) is given by Eq. 1 with an initial critical 1 the quantity M/ͱTH as a function of the scaling variable temperature Tc0 . Rigorously speaking, scaling is destroyed since both the for weak field ͑inset͒ and strong field ͑main part͒. We used ␭␪ here the maximal value ␪ ϭ1. In the strong field region, the concentration c and the shifted eigenvalue 1 depend ex- 1 plicitly on H and T. However, at strong field the correction deviation from clean-sample scaling behavior is negligibly term in Eq. ͑21͒ becomes negligibly small. This implies a small for all three values of the strong magnetic field, in complete agreement with our results. In the weak field re- restoration of the crossing point. Indeed, at temperature T0* the magnetization M*ϭM(T*,H) assumes the form gion, the scaling functions for three different fields can hardly be distinguished. This means that scaling is undoubt- 1 H⌽ edly valid in a neighborhood of the crossing temperature. At M*ϭM*ͩ 1Ϫ ͪ , ͑22͒ 0 1ϩ␩ HϩH* the same time the scaling function differs from its form in a clean sample ͑1͒ by a multiplicative constant ͑see the paren- ␩Ϫ1ϭ ␲ 2 Ј ␪ ⌽ ͑ ͒͒ with 2 L Hc2Tc0 1 / 0 . Therefore in the entire thesis in Eq. 16 . Note that scaling in the weak field region Low Temp. Phys. 27 (99–10), September–October 2001 Braverman et al. 751

This paper is devoted to the memory of Lev Shubnikov, whose outstanding contributions to many branches of low- temperature physics have shaped its development ever since.

This work was supported by the MINERVA Foundation ͑G. B.͒, by grants from Israel Academy of Science ‘‘Mesos- copic effects in type II superconductors with short-range pin- ning inhomogeneities’’ ͑S. G.͒, and ‘‘Center of Excellence’’ ͑Y. A.͒, and by a DIP grant for German-Israel collaboration ͑Y. A.͒. We would like to thank E. Zeldov, who drew our attention to the paper of van der Beek et al.,12 Z. Tesanovic´ for helpful discussions, and P. H. Kes, who advised us about ͱ FIG. 1. The quantity M/ HT as a function of the scaling variable x.The some parameters of the experimental setup of Ref. 12. ␮ ϭ dashed, dotted, and dot-dashed lines correspond to 0H 3, 4, and 5 T ͑ ͒ ␮ ϭ ͑ ͒ main figure and to 0H 0.02, 0.1, and 0.2 T inset . The solid line cor- *E-mail: [email protected] ͑ ͒ ͒ responds to the clean-sample scaling function strong field region only . 1 However, the field could not be very strong: if the order parameter is close to zero, vortices could collapse onto a single defect.16 But in the case Ӷ Ͻ H⌽ Hc2(T) there exists a finite region of strong fields H⌽ H Ͻ ͑which was experimentally established͒ is less pronounced Hc2(T) where the distribution of the vortices does not feel the presence than that in the strong field region. Apparently, the reason is of defects. that the experimental data are fitted to account for the clean- sample scaling function. Nevertheless, if we identify the fit- 1 P. H. Kes, C. J. van der Beek, M. P. Maley, M. E. McHenry, D. A. Huse, ted temperature 82.6 K ͑found in Ref. 12 in the weak field M. J. V. Menken, and A. A. Menovsky, Phys. Rev. Lett. 67,2382͑1991͒. ͒ ϭ 2 region with the renormalized critical temperature Tc Tc0 Q. Li, M. Suenaga, J. Gohng, D. K. Finnemore, T. Hikata, and K. Sato, Ϫ␦ Phys. Rev. B 46, 5857 ͑1992͒. Tc , and the fitted critical temperature 84.2 K in the 3 ´ 12 A. A. Abrikosov, Zh. Eksp. Teor. Fiz. 5, 1442 ͑1957͓͒Sov. Phys. JETP 27, strong field region with Tc0 , then, even within such a 1174 ͑1957͔͒. ␪ Ϸ ␪ 4 rough approximation, we obtain 1 0.5. Recalling that 1 E. Brezin, A. Fujita, and S. Hikami, Phys. Rev. Lett. 65, 1949 ͑1990͒. should be positive and less than unity, the above result 5 R. Ikeda, T. Ohmi, and T. Tsuneto, J. Phys. Soc. Jpn. 60, 1051 ͑1991͒. 6 Z. Tesanovic´, Phys. Rev. B 44, 12635 ͑1991͒. strongly supports the applicability of our theory to the perti- 7 12 D. Li and B. Rosenstein, Phys. Rev. B 60, 9704 ͑1999͒. nent experiment. 8 Z. Tesanovic´ and L. Xing, Phys. Rev. Lett. 67, 2729 ͑1991͒. 9 Z. Tesanovic´, Lei Xing, L. Bulaevskii, Q. Li, and M. Suenaga, Phys. Rev. ͑ ͒ 4. SUMMARY Lett. 69, 3563 1992 . 10 Z. Tesanovich and A. V. Andreev, Phys. Rev. B 49, 4064 ͑1994͒. In summary, we have calculated the magnetization of an 11 L. Civale, A. D. Marwick, M. W. McElfresh, T. K. Worthington, A. P. Malozemoff, F. H. Holtzberg, J. R. Thompson, and M. A. Kirk, Phys. Rev. irradiated superconductor below the mean-field transition Lett. 65, 1164 ͑1990͒; L. Civale, A. D. Marwick, T. K. Worthington, M. A. line Hc2(T), using the approach developed in Refs. 8–10 Kirk, J. R. Thompson, L. Krusin-Elbaum, Y. Sun, J. R. Clem, and F. H. and 15. It was shown that, from a rigorous point of view, Holtzberg, ibid. 67, 648 ͑1991͒. 12 disordered short-range defects are expected to destroy the C. J. van der Beek, M. Konczykowski, T. W. Li, P. H. Kes, and W. Benoit, Phys. Rev. B 54, R792 ͑1996͒. scaling behavior and prevent the existence of crossing point 13 I. B. Khalfin and B. Ya. Shapiro, Physica C 202, 393 ͑1992͒. in both regions of weak and strong magnetic fields ͑with 14 A. I. Buzdin, Phys. Rev. B 47, 11416 ͑1993͒. 15 ͑ ͒ respect to matching field H⌽͒. And yet, within the frame- Z. Tesanovic´ and I. F. Herbut, Phys. Rev. B 50, 10389 1994 . 16 G. M. Braverman, S. A. Gredeskul, and Y. Avishai, Phys. Rev. B 57, work of the parameters which were employed in the experi- ͑ ͒ 12 13899 1998 . mental work, the deviation from scaling behavior appears 17 G. M. Braverman, S. A. Gredeskul, and Y. Avishai, Phys. Rev. B 59, to be negligibly small, and crossing points exist in both field 12039 ͑1999͒. regions, in complete agreement with the experimental find- 18 I. M. Lifshits, Nuovo Cimento 3, Suppl., No. 4 ͑1956͒. 19 ings. The two fitting critical temperatures introduced in S. A. Gredeskul, M. Zusman, Y. Avishai, and M. Ya. Azbel’, Phys. Rep. 288, 223 ͑1997͒. Ref. 12 for the strong and weak field regions correspond, in our formalism, to the initial and renormalized critical This article was published in English in the original Russian journal. Repro- temperatures. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Commensurate vortex lattices in thin vanadium films and in VÕSi superconducting superlattices N. Ya. Fogel* and E. I. Buchstab

Department of Physics, Solid State Institute Technion, Haifa 32100, Israel V. G. Cherkasova, O. I. Yuzephovich, and M. Yu. Mikhailov

B. Verkin Institute for Low Temperature Physics and Engineering, National Academy of Sciences of Ukraine, pr. Lenina 47, 61103 Kharkov, Ukraine A. N. Stetzenko

Kharkov State Polytechnical University, ul. Frunze 21, 61002 Kharkov, Ukraine ͑Submitted March 21, 2001͒ Fiz. Nizk. Temp. 27, 1019–1029 ͑September–October 2001͒ The features of the Shubnikov phase in thin films of type-II superconductors are investigated in the case when the magnetic field is parallel to the surface of the film. Measurements of the nonmonotonic dependence of the critical current Ic on the magnetic field Hʈ reveal commensurate vortex lattices with different numbers of vortex chains in the film. It is proved experimentally that in homogeneous films the commensurability effect between the vortex lattice parameter and the thickness of the film can be observed only for the ideal state of the film boundaries, admitting the formation of an infinite lattice of the vortices and their images. Worsening of the film surface smoothness and of the plane-parallel precision of the two surfaces of the film leads to vanishing of the oscillations of Ic and to a sharp decrease of the critical current. The lock-in transition due to the influence of the surface barrier is observed in the films for the first time. It is found that a thin-film layered sample exhibits an interplay between two types of commensurability effects: with the period of the superstructure and with the total thickness of the sample. The H – T phase diagram is considered for a homogeneous film in a parallel magnetic field. © 2001 American Institute of Physics. ͓DOI: ͓DOI: 10.1063/1.1401184͔

Recently, commensurability effects in type-II supercon- perconductors was first studied by Guyon’s and Martinolli’s ductors ͑meaning commensurability of the lattice parameter groups.3,4 In one of the cases the objects of study were arti- a0 of the vortex lattice and the period of the regular structure ficial periodic superlattices consisting of a PbBi alloy with a of the inhomogeneities͒ have become a focus of research periodic modulation of the concentration of the components, attention as substantial progress has been achieved in the while in the other case they were thin films with an artifi- creation of nanostructures and periodic systems of pinning cially created periodic modulation of the thickness. In the the centers at extremely small length scales.1 Of no less interest aforementioned superlattices the curves of the critical current for the study of the commensurability effect are natural ma- Ic as a function of H exhibit oscillations when the magnetic terials with regular periodic inhomogeneities. These materi- field H is parallel to the plane of the layers. In the films with als include transition-metal dichalcogenides and high- modulated thickness the oscillations of Ic were observed in a temperature metal oxides. Many materials of this type— magnetic field perpendicular to the layers. In both cases the superconducting superlattices of different kinds—have also oscillations were observed for a mutually perpendicular ori- been created artificially. In the case when the intervortex entation of the vortices and the modulation direction. At cer- distance is equal to or a multiple of the period of the super- tain values of the magnetic field the period of the vortex structure of inhomogeneities, all of the vortices are pinned in lattice turns out to be commensurate4 with the period of the the pinning potential. This situation, as a rule, corresponds to modulation potential, and maxima of Ic appear namely at the maximum critical current Ic , and the corresponding vor- these fields. Later the effect under discussion was also stud- tex structure is called commensurate. In the incommensurate ied in other layered systems: artificial superlattices of differ- 5–12 phase, if no other types of pinning centers are present, the ent types and high-Tc superconductors. A theoretical ex- critical current should be equal to zero.2 Such an ideal struc- planation for the results mentioned can be found in Refs. 2, ture of a superconductor is not achievable in practice, and the 13, and 14. The oscillatory dependence of the critical current critical current is nonzero even in the incommensurate phase. Ic and resistance R as functions of Hʈ has also been observed Experimentally one observes oscillatory dependence of the in the situation when, because of strong intrinsic pinning, 3–8 9,10 5,6,15 critical current Ic , magnetic moment M, and energy only commensurate vortex lattices can exist. In that case 11,12 ͑ dissipation as functions of magnetic field. the maxima of Ic and minima of R) correspond to stable As far as we know, the commensurability effect in su- states of the commensurate vortex lattice or to phase transi-

1063-777X/2001/27(9–10)/8/$20.00752 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 Fogel et al. 753 tions between lattices with different orders of commensura- infinite regular lattice of images. Furthermore, such a lattice bility.5,6 cannot be formed if the surface has noticeable roughness. An interesting cycle of studies was done on supercon- The latter, as we know, suppresses the effectiveness of the ducting films with an artificial two-dimensional lattice of Bean–Livingston surface barrier. ͑ submicron pores or ferromagnetic particles see, e.g., Refs. It should be noted that the oscillations of Ic and M also 16–18͒. In those experiments the configuration, size of the take place in multilayered films of finite thickness.12,25,26,28 elements, and periodicity of the artificial lattice were varied, However those data were obtained for Nb/Cu and Nb/Ti su- and the critical current, magnetization, magnetic relaxation, perlattices with a small anisotropy parameter (␥ and current–voltage characteristics were measured. The field ϭ1.2– 1.8),12,28 which corresponds to a weak periodic po- dependences of all these parameters exhibited distinct fea- tential that has little influence on the properties of the film. tures in fields corresponding to commensurability conditions. Indeed, for homogeneous films of different materials and Of particular interest is the case in which the commen- Nb/Cu and Nb/Ti layered films there is no qualitative differ- surability effect is manifested in superconducting samples in ence in the behavior of Ic(Hʈ). In the case of thin-film su- which no periodic pinning potential is present. We are refer- perlattices there is only a renormalization of the values of the ring to homogeneous superconducting thin films, for which commensurability fields. This question will be discussed in oscillatory Ic(H) curves in a parallel magnetic field have more detail below. also been observed.19–24 In films whose thickness satisfies Our proposed interpretation for homogeneous films has the conditions ␭ϾdϾ␰ ͑␭ is the penetration depth of the found support in the paper by Carneiro,27 where the equilib- magnetic field, d is the thickness of the film, and ␰ is the rium configuration of the vortex lines in films was deter- coherence length͒, the transition from the Meissner to the mined numerically by the Monte Carlo method on the basis mixed state starts with the penetration of a single vortex of London theory with allowance for an infinite system of chain. As the external magnetic field is increased, the density image vortices. of vortex lines in the chain grows, and at a certain critical Since, in our view, the influence of the surface barrier is density an instability arises, causing the single row of vorti- exceedingly important and the commensurability effect in ces to split into two. Further increase of the field gives rise to homogeneous thin films can be observed only when the film a vortex lattice consisting of three chains, and so on. Owing has ideal boundaries, we undertook experiments expressly to the interaction of the vortices with the boundaries of the designed to permit unambiguous confirmation of this hypoth- film, a symmetric arrangement of the vortex chains with re- esis as to the origin of the oscillations of Ic in films in a spect to the two surfaces of the film corresponds to the most parallel magnetic field. stable state of the vortex lattice. When the distances between In addition, we obtained data on a superlattice sample of vortex chains are commensurate with the total thickness of finite thickness with larger anisotropy than in the multilay- the sample, extrema of Ic are observed. This is the basic ered films investigated previously. A comparison of these explanation offered for the oscillatory dependence of the film data with the results for homogeneous films leads to some parameters as functions of Hʈ in Refs. 12, 25, and 26. In interesting conclusions. On the one hand, in layered films it Refs. 22 and 23 a somewhat different interpretation was pro- is possible to have commensurability effects simultaneously posed. The interaction of the vortices with the boundaries of with respect to the total thickness of the sample and the the superconductor as a rule is treated by taking the image period of the superstructure. On the other hand, because of vortices into account. In the case of a thin film the system of the competing influence of the two different commensurabil- vortices and their images form an infinite lattice.27 The pe- ity effects an additional selection rule arises, so that the ex- riod of this lattice along the direction orthogonal to the film trema corresponding to the successive values of n do not all is determined by the thickness of the film and the number of appear. vortex rows, and the lattice period along the film is deter- The H – T phase diagram for a homogeneous supercon- mined by the external magnetic field. These two periods co- ducting film in a parallel magnetic field, is also discussed. incide or are multiples of each other only at certain values of 22,23 the magnetic field: SAMPLES AND EXPERIMENTAL TECHNIQUE As the objects of study we chose homogeneous vana- ϭ͑ 2)⌽ ͒ 2 ϭ ͑ ͒ Hn n 0 /2d , n 1,2,3,... 1 dium films. These films were condensed by electron-beam evaporation of the metal in a vacuum chamber at residual gas Here n is the order of commensurability or the number of pressures of ϳ10Ϫ6 Torr. The sharp-focus electron gun used ⌽ vortex rows in the film, and 0 is the magnetic flux quan- had a power of 5 kW. The material used for the deposition tum. Thus in order for oscillatory effects to appear it is im- was VE´ L-2 ͑99.9% vanadium͒. portant to have commensurability in the two-dimensional The substrates used were especially smooth glass. The ͑2D͒ lattice of vortices and image vortices. Under the condi- films were deposited on substrates heated to 300 °C. The rate tion of commensurability this 2D lattice has high symmetry, of deposition of the vanadium was held constant at a value and the lattice of the real vortices inside the film corresponds from 40 to 60 Å/s. Together with the correctly chosen sub- to an Abrikosov vortex lattice ͑for nϾ1͒. For the formation strate temperature, the electron-beam evaporation method of a regular periodic lattice of vortices and their images it is made it possible to obtain fine-grained films with an average necessary that the surfaces of the film be smooth and parallel crystallite size of 150–200 Å. We did not observe texture in to each other. Even a slight misorientation between the the films. Studies on a scanning electron microscope did not vortex-reflecting surfaces will preclude the formation of an reveal any noticeable flaws of the surface relief of the film. 754 Low Temp. Phys. 27 (99–10), September–October 2001 Fogel et al.

The samples were deposited through special heat- resistant masks of molybdenum foil 0.3 mm thick, prepared by an eletroerosion method. The distance between the mask and substrate was 0.1–0.3 mm. The geometry of the depos- ited samples facilitated the use of the four-probe method for studying their electrical characteristics. The characteristic di- mensions of the samples were as follows: distance between potential contacts 5 mm, width 0.5 mm. The thickness of the films during deposition was deter- mined by both a quartz oscillator and from the deposition time at a known rate. The thickness of the vanadium films was varied in the range 600–1600 Å. After preparation of the samples their thickness was monitored by the Talansky method to an accuracy of Ϯ20 Å. Besides the films, we also investigated V/Si superlattices FIG. 1. Dependence of the critical current Ic on the parallel magnetic field ϭ ͓ ͔ ͑ ͒ formed by the successive programmed deposition of vana- Hʈ for sample No. 1 (d 750 Å) at different temperatures T K : 1.980 1 , 2.384 ͑2͒, 2.749 ͑3͒, 3.125 ͑4͒, 3.368 ͑5͒, 3.618 ͑6͒, 3.819 ͑7͒, and 3.989 ͑8͒. dium and silicon layers on a fluorophlogopite substrate by the method of dc magnetron sputtering in an argon medium. The working pressure of the argon in the vacuum chamber not shift in field as the temperature is varied. It has been Ϫ Ϫ was 3ϫ10 3 Torr. The initial vacuum was 10 6 Torr or bet- shown previously29 that the field H* corresponds to either ter. The substrate temperature was maintained at 100 °C. The the first critical field Hc1 or to the surface barrier field Hs Ͼ deposition, as in the case of the thin films, was done through Hc1 , depending on the state of the film surface. As we a special heat-resistant mask with a specified geometry. The have said, penetration of vortices into the film in the field of thicknesses of the layers were determined by means of quartz the surface barrier occurs in the case of smooth boundaries of sensors from the masses of the deposited substances. The the film. The experiments done demonstrate a clear correla- error in the determination of the layer thicknesses was 1–2 tion: if the field H* is equal to the first critical field30 Å. The number of bilayers of the superlattice was 10. An 2⌽ ␨d electron microscope study ‘‘in transmission’’ showed that the ϭ 0 ͫ ϩ ͬ ͑ ͒ Hc1 2 ln 0.081 , 2 Si films are amorphous, while the V film is fine-grained poly- ␲d ␲␰͑T͒ crystalline, with a grain size of the order of 100 Å. 1͒ ͑ ␨ then there are no oscillations of Ic(Hʈ) Ref. 29; here Measurements of the superconducting characteristics ϭ ͒ 1.78 is Euler’s constant ;ifH* is equal to Hs , then the were done on an apparatus equipped with a superconducting Ic(Hʈ) curves are nonmonotonic, like those shown in Fig. 1. solenoid. The samples were placed in liquid helium, and no These data attest to the important role of the surface barrier heating was observed in the interval of currents used in the in the appearance of oscillation effects. experiments. The stabilization of the temperature at the cho- Oscillations of the critical current are observed not only sen point and the accuracy of its measurement were 0.003 K in a parallel field but also in inclined magnetic fields in a or better. To change the orientation of the films in the exter- certain region of angles near the parallel orientation. For nal magnetic field the samples were placed on a rod relatively small misorientations between the direction of the equipped with a special rotating device. The samples on the external magnetic field and the surface of the film the mini- rotating table could be arranged in two different ways: in the ͑ ͒ mum of Ic on these curves coincides Fig. 2 . However, start- first case the angle between the transport current and the magnetic field direction remained unchanged at 90° during the rotation, while in the second case changing the angle between the plane of the film and the magnetic field simul- taneously changed the angle between H and I as well. The parallel orientation was set according to the resistivity mini- mum to an accuracy of 0.1° or better. The critical current Ic was determined from the current–voltage characteristics ac- cording to the points at which the voltage reached a fixed value of 1 ␮V.

EXPERIMENTAL RESULTS AND DISCUSSION

Figure 1 shows the dependence of the critical current Ic on the parallel magnetic field Hʈ for vanadium sample No. 1 (dϭ750 Å) at various temperatures. The curves exhibit two features: 1͒ at magnetic fields lower than H*, which in- creases with decreasing temperature ͑its values are indicated FIG. 2. Dependence of the critical current Ic on the magnetic field H for ͒ sample No. 3 (dϭ1590 Å) at different angles ␪ between the direction of the by arrows in Fig. 1 the value of Ic remains practically un- external magnetic field and the plane of the sample: ␪ϭ0°. ͑1͒,5°͑2͒,8° changed as the field is increased; at the field H* the value of ͑3͒, and 11° ͑4͒. The angle ␪ is measured from the parallel orientation; T ͒ ϭ Ic starts to decrease sharply; 2 Ic has a minimum that does 3.614 K. Low Temp. Phys. 27 (99–10), September–October 2001 Fogel et al. 755

FIG. 4. Dependence of the critical current Ic on the reduced magnetic field FIG. 3. Dependence of the critical current Ic on the parallel magnetic field hϭH/Hʈ(T) for sample No. 3 at different temperatures. Hʈ for two different mutual orientations of the external magnetic field and the transport current for sample No. 3: IЌH ͑1͒, IʈH ͑2͒. Tϭ3.614 K.

The curves of I as a function of the reduced magnetic ␪ ͑ Ϸ ͒ c ing at a certain critical angle c for sample No. 3 it is 9° field hϭH/Hʈ(T) obtained at different temperatures form a the minimum vanishes, and the Ic(H) curves become mono- cigar-shaped family of curves, as follows from Figs. 4 and 5. tonic. The existence of a critical angle for observation of The values of Ic on those parts of the curves that bound the oscillations in the case of thin-film layered samples has also cigar coincide for different temperatures, i.e., there is no been reported previously in a number of papers ͑see, e.g., temperature dependence of Ic in a certain region of reduced Refs. 12 and 32͒. A more preferable explanation for the pres- fields and temperatures. In the region of fields h correspond- 33 ence of a critical angle is in terms of the lock-in transition ing to the interior of the cigar, where the minima and proposed in Ref. 32. Such a transition was predicted by maxima of Ic are located, one observes the usual type of Feinberg and Villard33 for layered superconductors with temperature dependence of Ic . For the region inside the ci- strong intrinsic pinning. The essence of this phenomenon is gar ͑and only for that region͒, as is shown in Fig. 6, there is that in a field inclined at small angles ␪ the minimum of the a characteristic hysteresis of Ic . Hysteresis in the neighbor- free energy corresponds to vortices oriented along extended hood of the fields Hn was also reported in Ref. 19. planar inhomogeneities ͑including the boundaries of the film͒ As we have said, the minima on the Ic(Hʈ) curves and not along the field. A difference appears between the should correspond to the conditions of commensurability be- directions of the magnetic induction in the sample and the tween the intervortex distance and the film thickness. The external magnetic field; this difference vanishes only for corresponding fields for a homogeneous film are determined ␪Ͼ␪ c . The authors of the cited papers did not rule out that in slightly inclined fields the state of the vortex system does not completely correspond to the picture described in Ref. 32. An alternative considered is a system of stepwise- inclined vortices, in which the main orientation coincides with the parallel one and those parts of the vortex lines which are parallel to the layers are connected together by so-called kinks oriented along the normal to the layers. Since our experiments have been done on homogeneous films, where stepwise inclined vortices cannot appear, their results permit us to state that for the first time a lock-in transition is observed which is due to the influence of a sur- face barrier rather than a periodic pinning potential. It follows from Fig. 3 that the presence or absence of minima in a parallel field is critically dependent on the ge- ometry of the experiment. The minimum is observed in the case when the vectors of the magnetic field and current, which lie in a plane parallel to the surface of the film, are mutually perpendicular ͑geometry 1 in Fig. 3͒. Here a Lorentz force arises which causes the vortices to move in the direction perpendicular to the surface of the film, overcom- ing the strong surface barrier. In the forceless configuration ͑geometry 2 in Fig. 3͒ there are no such features on the Ic(Hʈ) curves. Thus the surface barrier again emerges as a FIG. 5. Dependence of the critical current Ic on the reduced magnetic field phenomenon directly involved in the origin of the oscilla- hϭH/Hʈ(T) at different temperatures for sample No. 4 (dϭ1590 Å) before ͑ ͒ ͑ ͒ ϭ tions. 1 and after 2 etching; t T/Tc . 756 Low Temp. Phys. 27 (99–10), September–October 2001 Fogel et al.

FIG. 6. Dependence of the critical current Ic on the reduced magnetic field hϭH/Hʈ(T) for sample No. 3 for increasing and decreasing magnetic field; FIG. 7. Dependence of the critical current Ic on the parallel magnetic field Tϭ3.614 K. Hʈ for a 240/30 V/Si superlattice. The solid arrows indicate the calculated values of Hn , and the dotted arrow the calculated value of HN . by formula ͑1͒. If the film is a layered system, then formula In multilayer superlattices in which the boundary effects ͑1͒ transforms into the following form with allowance for the can be neglected, the fields HN at which the vortex lattice anisotropy parameter ␥:26 parameter is commensurate with the period of the superstruc- 34 ϭ 2)⌽ ␥ 2 ϭ ͑ ͒ ture of the superlattice are determined by the expression Hn n 0/2 d , n 1,2,3,... 3 H ϭ)⌽ /2N2s2␥. ͑4͒ The experimental data obtained in the present study are N 0 in good agreement with formula ͑1͒. For example, for sample Here N is the order of commensurability for the effect under ϭ ϭ No.1(d 750 Å) the field Hmin 3.1 kOe, and a calculation study. The ratio of the fields HN and Hn for the two type of ͑ ͒ ϭ ͑ ͒ according to formula 1 gives the value dcalc 760 Å for commensurability effect, as can be seen from formulas 3 ϭ ϭ ͑ ͒ n 1. For sample No. 2 (d 780 Å) one has Hmin and 4 , does not depend on any of the parameters of the ϭ ϭ 3.25 kOe and dcalc 740 Å. For the identical samples Nos. system except the number of bilayers: ϭ ϭ 3and4(d 1590 Å) the field Hmin 2.6 kOe, and the calcu- 2 HN m lated value of d for nϭ2 is 1660 Å. Good agreement with ϭ . ͑5͒ H n2N2 formula ͑1͒ has also been noted19 for homogeneous films of n the alloy PbIn with thicknesses dϭ650– 3500 Å.2͒ It is easy It follows from formula ͑5͒ that it is possible to observe to see that for Nb/Cu layered films the position of the a large number of extrema of Ic . We recall that upon com- 3͒ minima on the M(Hʈ) curves correspond to the calculated mensurability of the vortex lattice parameter a0 with the total ͑ ͒ fields Hn found using formula 3 . Good agreement with this thickness of the film one should observe minima of the criti- formula has also been observed in the case of Nb/Ti thin-film cal current Ic , while commensurability with s corresponds to 28 superlattices. maxima of Ic . It is of interest to compare the data for homogeneous It follows from an analysis of Fig. 7 that the first mini- ϭ thin films and films of the same material with a periodic mum of Ic for Hʈ 1.17 kOe corresponds to the calculated ϭ ϭ system of planar inhomogeneities. In the case of layered value of Hn at n 5, while the second at Hʈ 3.8 kOe corre- thin-film samples with sufficiently large values of the anisot- sponds to the value nϭ9. The minima with commensurabil- ropy parameter ␥ an ‘‘interference’’ between the two differ- ity orders nϭ1 – 4 cannot be observed because the fields ent commensurability effects can arise. The vortex lattice corresponding to them are less than or of the order of the parameter can be commensurate with both the sample thick- field at which the initial penetration of vortices into the film ness d and with the period s of the layered structure. The occurs. Such an explanation will of course not work for the available experimental results for Nb/Cu and Nb/Ti ͑with commensurability orders nϭ6 – 8, for which the extrema ␥ϭ ͒ values 1.2– 1.8 attest to the fact that in the case of a also fail to appear on the Ic(Hʈ) curves. It is important to comparatively weak modulation of the order parameter the note, however, that, since for the given sample mϭ10, it is presence of anisotropy leads to renormalization of the values only for nϭ5 and 9 that all the vortex chains can lie in the Si ͓ ͑ ͔͒ of Hn see formula 3 , but the surface barrier exerts the spacers, where their position is energetically most favorable. dominant influence on oscillation effects, as in the case of At nϭ9 all of the Si spacers in the sample are occupied, and homogeneous films. Here we present the results for a V/Si the distance between vortex rows is equal to s.Atnϭ5 the superlattice ͑vanadium layer thickness 240 Å, silicon spacer distance between vortex rows is equal to 2s. On the other ͒ thickness 30 Å with a comparatively large value of the an- hand, the position of the maximum of Ic at 4.7 kOe corre- isotropy parameter, ␥ϭ5.2. This sample consists of the same sponds, according to formula ͑4͒, to commensurability order ϭ ϭ ϭ number of bilayers, m 10, as the Nb/Cu and Nb/Ti samples N 1. The position of HN for N 2 coincides with that for 12,25,28 ϭ studied previously. The total thickness of the sample is n 5, and there a minimum of Ic is observed. A curious 2700 Å. The dependence of the critical current Ic on the picture emerges. Instead of the large number of oscillations parallel magnetic field for this sample is shown in Fig. 7. expected one observes only a few, and the reason for this is Low Temp. Phys. 27 (99–10), September–October 2001 Fogel et al. 757

apparently the additional selection rule that arises in the case Thus, from the last two experiments and also from the of a competition between two different commensurability ef- correlation between the appearance of the minimum of Ic fects. Of the set of vortex lattices that can exist in homoge- and the penetration of vortices into the film in the surface ͑ neous films, in a superlattice sample of the S/I type Sisa barrier field Hs it follows unambiguously that commensura- superconductor and I is a semiconductor͒ only those which bility between the film thickness and the interchain distance correspond to an ordered arrangement of vortex rows in the in the vortex lattice is by itself insufficient for observation of semiconductor spacer layers are realized. In relatively weak oscillation effects. On the other hand, the same experiments fields the dominant effect is that of commensurability with give evidence in favor of the previously stated hypothesis ͑ the total thickness of the film we recall that the fields Hn for that an important role is played by the formation of a regular ϭ ϭ n 5 and HN for N 2 are numerically equal, but at that lattice of vortices and their images and the conditions of ͒ value of the field a minimum of Ic is observed . In the case commensurability in that lattice. of high fields both commensurability effects are manifested: Having proved experimentally that the presence of a lat- ϭ ϭ a minimum of Ic at Hn 3.8 kOe (n 9) and a maximum of tice of vortices and vortex images vortices is critically im- ϭ ϭ Ic at HN 4.7 kOe (N 1). In both cases the distance be- portant for observation of the commensurability effect, we tween vortex chains is the same, equal to s, but the number can repeat the previous arguments22,23 concerning the nature of vortices in the chains is different. The appearance of the of the oscillatory critical current in thin-film samples. It can maximum after the minimum of Ic is explained by the fact be assumed that in accordance with the general concepts Ͼ that in fields Hʈ Hn the depth of the potential well for the about phase transitions between commensurate and incom- vortices increases.32,35 However, the theory does not deter- mensurate phases,37,38 in the neighborhood of the points of mine the field values up to which this occurs. For the given commensurability ͑i.e., in a certain interval of variation of sample the maximum at 4.7 kOe corresponds exactly to the the vortex lattice parameter or the external field H) there condition of commensurability between the superlattice pe- exists a commensurate ͑C͒ phase. In the C phase the vortices riod and a0 . It follows from the experiment under discussion behave in the usual way, and the critical current is deter- that in a superlattice of finite thickness the character of the mined by the mechanism of interaction of the vortices with oscillation effects at large values of the parameter ␥ will the boundaries of the film. According to the calculations of differ considerably from those for homogeneous films. Shmidt,35 in this region one should observe a minimum of Let us now return to the case of homogeneous films and the critical current, corresponding to the entry of a mono- present the results of direct experiments which indicate that layer of vortices, and a maximum due to fact that the depth the smoothness and plane-parallel precision of the two sur- of the potential well for the chain of vortices increases with faces of the films are the main factors determining the ap- increasing field. Of course, in this case the critical current 35 pearance of the oscillatory dependence of Ic(Hʈ). The sur- should depend on temperature. Just such behavior is ob- ͑ ͒ face of sample No. 4 after the measurements of the Ic(Hʈ) served inside the region bounded by the ‘‘cigar’’ see Fig. 4 . curves ͑the series of curves 1 in Fig. 5͒ was subjected to For each system there exists an incommensurability pa- chemical etching. After the etching its thickness had changed rameter ␦, at the critical value of which a transition should only slightly (⌬d/dϷ5%). An examination with a scanning occur to an incommensurate ͑I͒ phase. In the I phase, in electron microscope ͑resolution ϳ200 Å͒ showed that the accordance with the same concepts,37,38 one typically sees state of the surface had been modified by the chemical etch- the formation of static periodically repeating solitons ͑do- ing. Whereas before etching it had been practically smooth main walls͒. In our case these solitons or domain walls can ͑within the resolution of the device͒, after the etching there separate regions with two different commensurate lattices. were shallow ‘‘pits’’ of arbitrary shape, with dimensions of The critical current on the envelope of the ‘‘cigar’’ is ␮ 0.1–1 m. After the etching the values of Ic and the form of independent of temperature and can apparently be interpreted ͑ the Ic(Hʈ) curves were radically altered series of curves 2 in as the critical current due to the transition from the I phase to Fig. 5͒. The values of the critical current in the same interval the C phase, or as the current at which the solitons are de- of reduced temperatures and fields decreased by a factor of stroyed. Inside the C phase the critical current is determined ͑ 3–7 while Tc and Hcʈ changed insignificantly, by not more by the relationships governing the interaction of individual ͒ than 2–4% . The Ic(Hʈ) curves became monotonic: the ‘‘ci- vortices with the boundaries of the film. gar’’ vanished. The artificially created roughness of the The results of the above experiments allow us to state sample surface led to suppression of the surface barrier and, with certainty that the oscillations of the superconducting accordingly, to an overall decrease of the critical current. The parameters of homogeneous films in a parallel magnetic field features characteristic of the commensurability effect also are due to the commensurability effect occurring in the lat- vanished. tice of vortices and vortex images vortices arising in the case In the same deposition session as for sample No. 1 we of an ideal state of the boundaries of the sample. These re- also prepared a sample ͑No. 5͒ with variable thickness by the sults are consistent with the results of previous experiments technique proposed in Ref. 36. For this wedge-shaped which established that for observation of oscillations it is sample with a difference in thicknesses at the edges of ϳ7% necessary to have commensurability between the vortex lat- and a misorientation of the two surfaces amounting to only a tice parameter in the film and the total thickness of the few minutes of arc, the values of Ic at the same values of t sample, since the conditions of commensurability are the and h are smaller by a factor of 2–3.5 than for sample No. 1, same in both cases. For our hypothesis, however, one must ͑ and the Ic(Hʈ) curves are smooth the average thickness of impose more stringent requirements on the samples in order film No. 5 was close to the thickness of sample No. 1͒. to be able to observe the effects under discussion. These 758 Low Temp. Phys. 27 (99–10), September–October 2001 Fogel et al. conditions undoubtedly must be met in the case of homoge- neous samples. If the film has a layered structure, then one can observe oscillations of the physical parameters due to the commensurability between the interchain distance and the period of the superstructure, and therefore the state of the surface is not so important. Let us conclude with a discussion of the H – T phase diagram of a homogeneous superconducting film in a parallel magnetic field and the necessary conditions for observation of the commensurability effect due to the influence of the sample boundaries. One of the obvious conditions, a restric- tion on the film thickness, ␭ӷdӷ␰, has already been men- tioned above. It follows from this condition that the materials used for preparation of the films should have rather large values of the Ginzburg–Landau parameter ␬. This condition is met for transition metals, superlattices based on them, and certain superconducting alloys. The inequality dӷ␰(T) pre- supposes the possibility of penetration of one, two, etc. chains of vortices into the film in a parallel magnetic field. This inequality, which, with allowance for the critical thick- FIG. 8. H – T phase diagram for sample No. 1: M, S, and N are the Meiss- ϭͱ ␰ ͑ ͒ ϭ ness dc 5/2 (T) Ref. 39 , determines the condition under ner, mixed, and normal states. The horizontal lines with n 1,2,3 indicate which the homogeneous film undergoes a transition to a one- the positions of the commensurate phases, and the shading indicates the existence region of each of these phases. dimensional mixed state at the upper critical field, can be уͱ ␰ reduced to the less stringent restriction d 5/2 (T0). If this condition is not met, then the uniform distribution of the moment is six.12 In the case when measurements of the criti- order parameter will persist in the film all the way down to cal current are used to study the commensurate phases, one Tϭ0K.Ifdӷ␰(0), then the upper boundary for the exis- ordinarily sees fewer of them. To a large degree this is due to tence of the mixed state is determined, as in the case of a a purely ‘‘technical’’ limitation: at large values of the critical ϭ bulk superconductor, by the field Hc3 1.7Hc2 current density the samples can become overheated ͑e.g., in ϭ ⌽ ␲␰2 1.7 0 /(2 ). From the field ratio Hc3(0)/Hn sample No. 1 at tϷ0.5 the critical current density in low Ϸ0.2d2/͓n2␰2(0)͔ it follows that the maximum number of fields already exceeds 106 A/cm2͒. Measurements of the commensurate vortex phases with different n which can be magnetic moment, which are free of this limitation, can give observed at Tϭ0 in a given film is determined by the ratio more complete information about commensurate vortex lat- ␰ Ͻ ␰ d/ (0): nmax 0.47d/ (0). This definition of nmax is valid tices in thin films. for those film thicknesses for which the condition d/␰(0) р4 is met. If this inequality does not hold, then the com- This study was supported in part by the Israel Academy ͑ ͒ mensurate lattice with nϭ1atTϭ0 or at low temperatures of Sciences Grant 359/99 and the Center for Absorption in Science of the Ministry of Immigrant Absorption of the State cannot be detected experimentally, since the field Hc1 turns out to be less than the field of overheating of the Meissner of Israel. ϭ⌽ &␲␰ ϭ 35 state, Hs 0 /( d), at T 0. Accordingly, the num- * ber of commensurate vortex lattices will be equal to nmax E-mail: [email protected] or [email protected] ͒ Ϫ1. An example of the H – T phase diagram for a homoge- 1 Oscillations can be absent for a completely trivial reason as well ͑see, e.g., ͒ neous film is shown in Fig. 8. Ref. 31 . The critical current in that study was measured with such a large step in magnetic field that it was not possible to register oscillations. At nonzero temperatures the number of observable com- 2͒Sutton’s formula,19 which is based on a calculation of close packing of mensurate vortex phases depends on T, since the fields spheres, is practically the same as formula ͑1͒. 3͒ 25,26 bounding the existence region of the mixed state are tem- Some authors have held the opinion that the commensurability condi- tions correspond to maxima rather than minima of I and M, but appar- perature dependent. For this reason there exists a region near c ently this point of view has changed, since a more recent paper12 by some Tc in which one can observe a commensurate phase with n of the same authors spoke of minima. ϭ1, which at lower temperatures is ‘‘hidden’’ in the region of the Meissner state. As T is lowered, this phase vanishes, ϭ but the phase with n 2 becomes observable. At still lower 1 V. V. Moshchalkov, Physica C 332,1͑2000͒. temperatures one can observe two or more commensurate 2 S. E. Burkov and V. L. Pokrovsky, J. Low Temp. Phys. 44,423͑1981͒. phases. The larger the film thickness, the larger the number 3 H. Raffi, J. C. Renard, and E. Guyon, Solid State Commun. 11, 1679 of commensurate phases there are in the region of the mixed ͑1972͒. 4 O. Daldini, P. Martinolli, T. L. Olsen, and G. Berner, Phys. Rev. Lett. 22, state. However, there is a limit to the film thickness, above 218 ͑1974͒. which the neighboring commensurate phases turn out to be 5 N. Ya. Fogel, M. Yu. Mikhailov, Yu. B. Bomze, and O. I. Yuzephovich, located so close to each other in magnetic field that the ex- Phys. Rev. B 59, 3365 ͑1999͒. 6 istence regions of the individual phases begin to overlap. The N. Ya. Fogel, M. Yu. Mikhailov, Yu. B. Bomze, and O. I. Yuzephovich, Physica B 284–288,731͑2000͒. maximum number of commensurate vortex phases that have 7 P. R. Broussard and T. H. Geballe, Phys. Rev. B 37,68͑1988͒. been observed in thin films by measurement of the magnetic 8 C. Coccorese, C. Attanasio, L. V. Mercaldo, M. Salvato, L. Maritato, J. M. Low Temp. Phys. 27 (99–10), September–October 2001 Fogel et al. 759

Slaughter, C. M. Falco, S. L. Prischepa, and B. I. Ivlev, Phys. Rev. B 57, 25 S. H. Brongersma, E. Verweij, N. Y. Koeman, D. G. de Groot, R. Griessen, 7922 ͑1998͒. and B. I. Ivlev, Phys. Rev. Lett. 71, 2319 ͑1993͒. 9 M. Oussena, P. A. J. de Groot, R. Gagnon, and L. Taillefer, Phys. Rev. 26 S. H. Brongersma, E. Verweij, N. Y. Koeman, D. G. de Groot, and Lett. 72,3606͑1994͒. R. Griessen, Thin Solid Films 228, 201 ͑1993͒. 10 A. A. Zhukov, H. Kupfer, G. K. Perkins, A. D. Caplin, K. I. Kugel, A. L. 27 G. Carneiro, Phys. Rev. B 57, 6077 ͑1998͒. Rakhmanov, M. G. Mikheev, V. I. Voronkova, M. Klaser, and H. Wuhl, 28 P. Lobotka, I. Varva, R. Senderak, D. Machajdik, M. Jergel, S. Gazi, E. ͑ ͒ Phys. Rev. B 59, 11213 1999 . Rossel, M. Baert, Y. Bruynseraede, M. Forsthuber, and G. Hilscher, 11 C. Hunnekes, H. G. Bohn, W. Shilling, and H. Schulz, Phys. Rev. Lett. 72, Physica C 229, 231 ͑1994͒. 2271 ͑1994͒. 29 ͑ ͒ 12 V. G. Cherkasova and N. Ya. Fogel’, Fiz. Nizk. Temp. 15,383 1989 M. Ziese, P. Esquinazi, P. Wagner, H. Adrian, S. H. Brongersma, and R. ͓ ͑ ͔͒ ͑ ͒ Sov. J. Low Temp. Phys. 15,216 1989 . Griessen, Phys. Rev. B 53,8658 1996 . 30 ´ ͑ ͓͒ 13 ͑ ͒ A. A. Abrikosov, Zh. Eksp. Teor. Fiz. 47,720 1964 Sov. Phys. JETP 20, S. Ami and K. Maki, Prog. Theor. Phys. 53,1 1997 . ͑ ͔͒ 14 ͑ ͒ 480 1965 . S. E. Burkov and V. L. Pokrovskii, JETP Lett. 32, 263 1980 . 31 15 G. Stejic, A. Gurevich, E. Kadyrov, D. Christen, R. Joyat, and D. C. B. I. Ivlev, N. B. Kopnin, and V. L. Pokrovsky, J. Low Temp. Phys. 80, ͑ ͒ ͑ ͒ Larbalestier, Phys. Rev. B 49, 1274 1994 . 187 1990 . 32 16 E. Rossel, M. Van Bael, M. Baert, R. Jonckheere, V. V. Moshchalkov, and A. J. Vermeer, D. J. de Groot, N. J. Koeman, R. Griessen, and C. van ͑ ͒ Y. Bruynseraede, Phys. Rev. B 53, R2983 ͑1996͒. Hasendonck, Physica C 185–189,2345 1991 . 33 ͑ ͒ 17 M. Baert, V. V. Metlushko, C. D. Potter, E. Rossel, K. Temst, R. D. Feinberg and C. Villard, Phys. Rev. Lett. 65, 919 1990 . 34 ͑ ͒ Jonckheere, A. Gilabert, V. V. Moshchalkov, and Y. Bruynseraede, Physica B. I. Ivlev and N. B. Kopnin, J. Low Temp. Phys. 80, 161 1990 . 35 ´ C 235–240,2791͑1994͒. V. V. Shmidt, Zh. Eksp. Teor. Fiz. 57, 2095 ͑1969͓͒Sov. Phys. JETP 30, 18 M. J. Bael, K. Temst, L. V. Look, J. Bekaert, V. V. Moshchalkov, and 1137 ͑1970͔͒. 36 ´ Y. Bruynseraede, Physica B 284–288, 893 ͑2000͒. Yu. F. Komnik and E. I. Bukhshtab, Zh. Eksp. Teor. Fiz. 54,34͑1968͒ 19 J. Sutton, Proc. Phys. Soc. 67,791͑1966͒. ͓Sov. Phys. JETP 27,34͑1968͔͒. 20 T. Yamashita and Y. Onodera, Solid State Commun. 13, 1289 ͑1973͒. 37 P. Bak, Rep. Prog. Phys. 45, 587 ͑1982͒. 21 ͑ ͒ 38 ´ ͑ ͒ T. Yamashita and L. Rinderer, J. Low Temp. Phys. 24, 695 1976 . V. L. PokrovskiŽ and A. L. Talapov, Zh. Eksp. Teor. Fiz. 78,269 1980 22 N. Ya. Fogel and V. G. Cherkasova, Physica B 107, 291 ͑1981͒. ͓Sov. Phys. JETP 51, 134 ͑1980͒; Erratum, 51, 654 ͑1980͔͒. 23 N. Ya. Fogel’ and V. G. Cherkasova, Fiz. Nizk. Temp. 7,268͑1981͓͒Sov. 39 I. O. Kulik, Zh. E´ ksp. Teor. Fiz. 52,1632͑1967͓͒Sov. Phys. JETP 25, J. Low Temp. Phys. 7, 131 ͑1981͔͒. 1085 ͑1967͔͒. 24 L. P. Ichkitidze and V. I. Skobelkin, Fiz. Nizk. Temp., 7,117͑1981͓͒Sov. J. Low Temp. Phys. 7,58͑1981͔͒. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Triplet superconducting proximity effect in inhomogeneous magnetic materials A. Kadigrobov*

Department of Applied Physics, Chalmers University of Technology and Go¨teborg University, SE-412 96 Go¨teborg, Sweden and B. Verkin Institute for Low Temperature Physics and Engineering, National Academy of Science of Ukraine, 47 Lenin Ave., Kharkov, 61103, Ukraine R. I. Shekhter and M. Jonson

Department of Applied Physics, Chalmers University of Technology and Go¨teborg University, SE-412 96 Go¨teborg, Sweden ͑Submitted May 14, 2001͒ Fiz. Nizk. Temp. 27, 1030–1038 ͑September–October 2001͒ We show that quantum spin fluctuations in inhomogeneous conducting ferromagnets drastically affect the Andreev reflection of electrons and holes at a ferromagnet–superconductor interface. As a result, a strong long-range proximity effect appears, associated with electron–hole spin triplet correlations and persisting on a length scale typical for nonmagnetic materials but anomalously large for ferromagnets. For applications, an important consequence is that this long-range proximity effect permits the creation of superconducting quantum interference devices with magnetic Josephson junctions of anomalously large length. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1401185͔

⌬ ϭ␮ ͑␮ 1. INTRODUCTION controlling the spin splitting E BH B is the Bohr magneton͒ of the electron and hole energies in the hybrid. In recent years much attention has been paid to normal It becomes necessary to take this splitting into account in conductor-superconductor ͑N/S͒ structures ͑for a review, see, the case of a ferromagnet-superconductor structure, where e.g., Ref. 1͒. Transport of electric charge in such systems is the interaction of the quasiparticle spin with the magnetiza- much affected by the existence of an energy gap in the spec- tion is of an exchange character and hence can be extremely trum of elementary excitations in the superconductor. As a large. Conservation of the electron-hole symmetry ͑that is, result of the existence of the gap, electronic elementary ex- the symmetric positioning of their energies with respect to citations which freely propagate in the nonsuperconducting ͒ ⌬ ϭ the Fermi level causes an additional difference p I0 /vF material cannot penetrate into the superconductor to a suffi- ͑I is the exchange energy of the ferromagnet͒ in the mo- ␧͑ 0 cient distance if their energy measured from the Fermi menta of the electron and hole components of the hybrid that level ␧ ͒ is less than the superconductor energy gap ⌬.A F drastically decreases the penetration length L␧ by orders of correlated transfer of two electrons accompanied by their ͑ magnitude in this case the penetration length is LI pairing inside the superconductor is the only mechanism that 0 ϭͱបD/I ; Ref. 2͒. Such a shortening of the proximity effect provides a direct transmission of the charge into the super- 0 has been actually observed in magnetic materials.3–7 On the conducting condensate that is the ground state of the super- other hand, measurements carried out in recent works8–11 conductor. The above-mentioned two-electron transfer may be considered in terms of the conventional scattering scheme demonstrate a long-range proximity effect in magnetic mate- rials that is in an obvious contradiction with the general con- as a process of an electron-hole transformation of excitations 9 inside the normal conductor that takes place at the boundary siderations discussed above. It has been pointed out that with the superconductor. This scattering ͑which is known as spin triplet fluctuations in the electron-hole correlations Andreev reflection͒ couples the incident electron and the re- caused by the spin-orbit interaction and electron-impurity 12 ͑ ͒ flected hole in such a way that their spins are oriented in scattering cannot by two orders of magnitude explain the opposite directions and their energies ͑Ϯ␧͒ are symmetri- large effect observed in Refs. 8–11. cally positioned with respect to the Fermi energy ͑‘‘Andreev The main message of this paper is that in magnetically hybrid’’͒. Such a two-electron correlation which arises at the inhomogeneous materials ͑such as multidomain ferromag- boundary with the superconductor persists inside the normal nets ͑F͒, inhomogeneous ‘‘cryptoferromagnetic’’ states im- 13 conductor to a distance L␧ from the superconductor, L␧ posed by the superconductor, or F/S interfaces inducing ϭ ប ⌬ ͱប ⌬ ⌬ ϭ␧ ͒ 14 min( / p, D/ pvF), where p /vF , vF is the electronic spin-flip processes , strong quantum fluctuations Fermi velocity, and D is the diffusion coefficient. The de- of the electron and hole spins make the proximity effect less struction of the phase coherence arises due to the difference sensitive to the spin selection rule that applies to Andreev ␧ between the momenta of the electron ( /vF) and hole reflections. As a result, a strong long-range, spin-triplet prox- Ϫ␧ ( /vF) components in the Andreev hybrid. The other pe- imity effect in F/S structures persists on a length scale typical ͒ culiar feature of the Andreev hybrid, which is that the elec- for nonmagnetic materials.1 We estimate the conductance of tron and hole spins have opposite directions, makes for sen- such an F/S structure to be of the same order of magnitude as sitivity of the Andreev correlation to a magnetic field H the conductance measured in experiments.8–11 Additional ex-

1063-777X/2001/27(9–10)/7/$20.00760 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 Kadigrobov et al. 761

FIG. 2. Schematic view of an S/F structure with a magnetic domain wall at ϭ x 0, a distance LD from the S/F interface. Impurity scattering is assumed to occur out of the ballistic region ͑to the right of the vertical dashed line͒.

2. FORMULATION OF THE PROBLEM We consider the conductance of a ferromagnet- superconductor structure schematically shown in Fig. 2 for a FIG. 1. Sketch of the composite scattering of an electron impinging on the special case when magnetic spin scattering occurs within a interface between a magnetically inhomogeneous ferromagnet and a super- distance, LD , from the S/F interface that is shorter than the conductor: Andreev reflection at the F/S interface is accompanied by spin- electronic mean free path l ͑that is, L Ӷl ͒. This allows us splitting scattering in the region of inhomogeneity ͑for clarity here assumed 0 D 0 to be confined to a finite-width layer͒. As a result, the reflected hole is in a to consider the Andreev reflection at the interface to occur mixed state with both spin up and spin down components. with unit probability and to describe the magnetic spin scat- tering using the semiclassical Eilenberger equation,17 which can be readily solved in the ballistic transport regime. Proper boundary conditions for matching the solution of the Eilen- periments with intentionally introduced magnetic inhomoge- berger equation to the appropriate solution in the diffusive neities are needed to check the predicted effect quantita- part of the ferromagnet can also be formulated in this model tively. and used to solve the Usadel equation,18 which is the appro- A schematic illustration of Andreev reflection in the priate equation in the diffusive transport regime. In this way, presence of a magnetic inhomogeneity is presented in Fig. 1. the excess conductance of the F/S boundary can be calcu- For convenience we consider the inhomogeneous magnetiza- lated. tion of the material to be confined to a layer of finite width Solving the model problem described, we find that a new located close to the superconductor-ferromagnet interface. type of superconducting ordering, corresponding to the trip- Such a layer serves as a magnetic spin-splitter for the inci- let spin correlations dent electron ͑see Fig. 1͒. The composite scattering produced by the F/S boundary and the spin-splitter can then be sepa- i␧t F ͑ ␧͒ϭ ͵ ͗␺ ͑ ͒␺ ͑ ͒͘ ͩ ͪ ͑ ͒ rated into three scattering events: an incident electron with ␴,␴ r; ␴ r,t ␴ r,0 exp ប dt 1 spin up crosses the inhomogeneous magnetic layer at point A ͑ ␴ϭ↑ ↓͒ and splits up into a coherent mixture of spin up (e,↑) ͑see here , , is the source of the proximity effect at dis- tances of order L ӷL , L ϵL␧͉␧ϭ . Fig. 1͒ and spin down (e,↓) electronic states. These are sub- T l0 T kT ject to Andreev reflection at the F/S interface and are trans- The Hamiltonian describing the system is written as fol- formed into spin down (h,↓) and spin up (h,↑) hole states lows: ͑see the dashed lines in Fig. 1͒. These two states encounter ˆ ϭ ͵ ͕␺ϩ͑ ͒͑ 2 Ϫ ͑ ͒͒␺ ͑ ͒ the magnetic scatterer again at points B1 and B2 , respec- H dr ␣ r pˆ /2m eV r ␣ r tively, and experience further ‘‘spin-splitting.’’ The final re- ϩ ϩ sult of the composite scattering process is that the incoming ϩ⌬͑r͒␺↑ ͑r͒␺↓ ͑r͒ϩ⌬*͑r͒␺↓͑r͒␺↑͑r͖͒ electron is reflected in two hole channels, one with spin up ϩ ͑ ͒␴ ␺ϩ͑ ͒␺ ͑ ͒ ͑ ͒ (h,↑) and the other with spin down (h,↓). One of the re- h r ␣␤ ␣ r ␤ r 2 flected hole channels has the same spin orientation as the where V(r) is the electrical potential; ␴ˆ are Pauli matrices; incident electron. ␣,␤ϭ(↑,↓) and summation with respect to double indices is ϭ Taking an alternative point of view, one may consider assumed; h(r) I0e(r) is the inhomogeneous exchange en- the time-reversed process wherein a Cooper pair propagates ergy in the ferromagnet ͑e(r) is a unit vector along the mag- from the superconductor to the ferromagnet. Being in the netization at point r͒; the superconductor energy gap ⌬ and singlet state at the moment of injection, the pair is then scat- the ferromagnet exchange field h have nonzero values in tered into the triplet configuration by an inhomogeneously complementary space regions: ⌬Þ0, hϭ0 in the supercon- oriented magnetization in the ferromagnet. This singlet- ductor and hÞ0, ⌬ϭ0 in the ferromagnet. triplet scattering is effective if the length scale of the inho- We start with the 4ϫ4 correlation functions mogeneity is of the order of L , the separation between the I0 ˆ Ͻ͑ Ј͒ϭ ͗␺ˆ ϩ͑ ͒␺ˆ ͑ ͒͘ ˆ Ͼ͑ Ј͒ two Cooper pair electrons in the ferromagnet. G 1,1 i 1 1 :G 1,1 The above result of the two-channel magneto-Andreev ϭϪi͗␺ˆ ͑1͒␺ˆ ϩ͑1͒͘ ͑3͒ scattering implies that the electron-hole correlation has a ϩ ϩ ϩ contribution that is unaffected by the magnetic exchange en- where ⌿ˆ (1)ϭ(⌿↑ (1),⌿↓(1),⌿↓ (1),⌿↑(1)) is the ergy, which leads to a long-range ‘‘spin-triplet’’ proximity Nambu pseudo-spinor field ͑its variable is 1ϵ(r,t)͒. Using effect. the correlation functions Eq. ͑3͒ in the standard way ͑see, 762 Low Temp. Phys. 27 (99–10), September–October 2001 Kadigrobov et al. e.g., the review article͒,19 one constructs the 8ϫ8 Green’s In the ferromagnet at distances much greater than the mean ϫ ϫ ͑ ͒ ͑ ͒ function in the 2 2 Keldysh and 4 4 Nambu spaces as free path l0 the dirty limit the Eilenberger equation 8 follows: reduces to the Usadel equation for the symmetric part of the ˇ ϵ͗ ͘ ͑͗ ͘ R K Green’s function G gˇ ... denotes an average over the Gˆ Gˆ ͒ Gˇ ϭͩ ͪ ͑4͒ directions of the electron/hole momenta : 0ˆ Gˆ A ␣ ␣ ␣ ␣ G↑↑ F↑↓ G↑↓ F↑↑ R,A,K ␣ ␣ ␣ ␣ where Gˇ are retarded ͑R͒, advanced ͑A͒, and Keldysh ¯F G¯ ¯F G¯ ␣ ↓↑ ↓↓ ↓↓ ↓↑ (K)4ϫ4 matrix Green’s functions which include both the Gˇ ϵ ␣ ␣ ␣ ␣ . ͑11͒ ͩ G↓↑ F↓↓ G↓↓ F↑↓ ͪ ˆ ˆ singlet G␴,Ϫ␴ and triplet G␴,␴ components of the normal as ␣ ␣ ␣ ␣ well as anomalous Green’s functions: ¯F↑↑ G¯ ↑↓ ¯F↑↓ G¯ ↑↑ ˆ Rϭ⌰ Ϫ ͒ ˆ Ͼ ͒Ϫ ˆ Ͻ ͒͒ This equation reads as follows: G ͑t1 t1Ј ͑G ͑1,1Ј G ͑1,1Ј d dGˇ Gˆ AϭϪ⌰͑t Ϫt ͒͑Gˆ Ͼ͑1,1Ј͒ϪGˆ Ͻ͑1,1Ј͒͒ ͑5͒ ͩ ˇ ͪ ϩ͓͑␧␶ Ϫˇ ͒ ˇ ͔ϭ ͑ ͒ 1Ј 1 D G ˇ I , G 0, 12 dx dx 3 0 ˆ Kϭ ˆ Ͼ͑ ͒ϩ ˆ Ͻ͑ ͒ G G 1,1Ј G 1,1Ј . ˇ ϭ ␴ˆ  ␴  ␴ ͑ ͒ where I0 I0 0 ˆ 3 ˆ 3 ; when writing Eq. 12 we took տ Using Eqs. ͑2͒–͑5͒, one gets the Eilenberger equation for the into account the fact that at distances x l0 from the bound- matrix quasi-classical Green’s function in the Wigner repre- ary the ferromagnet is assumed to be homogeneous, with the ϭ sentation, magnetic moment parallel to the z axis (hz I0). Here and below we also assume all quantities in the structure to vary only along the x axis which is perpendicular to the F/S inter- gˇ ͑n,R;␧͒ϭ͑i/␲͒␶ˇ ͵ d␰Gˇ ͑p,R;␧͒, ͑6͒ 3 face. ϫ ␶ where the 8 8 matrix ˇ 3 is represented in a compact nota- ␶ ϭ␴ tion of the tensorial product of Pauli matrices as ˇ 3 ˇ 0 3. BOUNDARY CONDITIONS FOR THE USADEL EQUATION  ␴  ␴ ˇ ␧ ˆ z ˆ 0 ; the integrand G(p,R; ) is the Fourier transform In order to find boundary conditions for the Usadel equa- ͑ ͒ of the Green’s function of Eq. 4 with respect to the coor- tion in the diffusive region we solve the Eilenberger equation dinate and time differences; the space variable in the center- ͑8͒ at distances xՇbӶl ͑see Fig. 2͒, where the impurity ϭ ϩ Ј 0 of-mass system is R r r ; the vector n is the unit vector scattering term (ϰ␴ˇ ) is negligible ͑we assume the supercon- ␰ϭ 2 Ϫ␧ along the momentum p, and p /2m F . For reference ␰ ϭប ͉⌬͉Ӷ ͒ ␣ ductor to be pure as well; that is, 0 vF / l0 . In this we write out gˆ in full as follows: case Eq. ͑8͒ reduces to the following linear equation: ␣ ␣ ␣ ␣ g↑↑ f ↑↓ g↑↓ f ↑↑ d ϩ ␧␶ ϩ⌬ˇ ͒Ϫ ˇ ͒͒ ϭ ͑ ͒ ␣ ␣ ␣ ␣ ivFnx ˇg ͓͑ ˇ 3 ͑x h͑x , gˇ ͔ 0. 13 ¯f ↓↑ ¯g↓↓ ¯f ↓↓ ¯g↓↑ dx ␣ϵ ͑ ͒ gˇ ␣ ␣ ␣ ␣ 7 →Ϫϱ ͩ g↓↑ f ↓↓ g↓↓ f ↑↓ ͪ Deep inside the superconductor (x ) we use the conventional boundary conditions for the Eilenberger equa- ¯␣ ␣ ¯␣ ␣ f ↑↑ ¯g↑↓ f ↑↓ ¯g↑↑ tion ͑8͒: where ␣ϭR,A,K. The Eilenberger equation for gˇ (n,R;␧) ␧ ⌬ reads g␴ ␴ϭ , f ␴ Ϫ␴ϭ , , ͱ␧2Ϫ͉⌬͉2 , ͱ␧2Ϫ͉⌬͉2 ץ iv n ˇgϩ͓␧␶ˇ ϩ⌬ˇ ϩi␴ˇ Ϫhˇ ,gˇ ͔ϭ0, ͑8͒ ⌬ ␧ * R 3ץ F ¯f ␴ Ϫ␴ϭϪ , ¯gϪ␴ Ϫ␴ϭϪ , ͑14͒ , ͱ␧2Ϫ͉⌬͉2 , ͱ␧2Ϫ͉⌬͉2 where ͓Aˇ ,Bˇ ͔ϭAˇ Bˇ ϪBˇ Aˇ ; the impurity-induced self-energy ␴ˇ ϭ ϭ ϭ¯ ϭ in the Born approximation is g␴,Ϫ␴ ¯g␴,Ϫ␴ f ␴,␴ f ␴,␴ 0. ␲ dn As we consider the case that the transparency of the ␴ϭ ͵ ͑ ␧͒ ͑ ͒ ˇ ␲ ˇg n,R, , 9 superconductor-ferromagnet interface is equal to 1 ͑that is 2t0 4 the Andreev reflection takes place in the absence of the nor- ⌬ˇ t0 is the impurity scattering time; the pairing potential that mal reflection͒ the boundary condition at the F/S interface ϭϪ ͑ ͒ determines the electron-hole correlations and the operator hˇ x LD see Fig. 2 is the continuity of Green’s functions, ϭ ˇ ϩ ˇ that is hЌ hz that describes the effect of the inhomogeneous magnetic moment on the spins of electrons and holes can be Ϫ Ϫ ͒ϭ Ϫ ϩ ͒ ͑ ͒ gˇ ͑ LD 0 gˇ ͑ LD 0 . 15 written as We solve the ballistic linear Eilenberger equation ͑13͒ in ⌬ˇ ϭ␴  ␴  ⌬ ␴ Ϫ⌬␴ ͒ ␴ ϭ␴ Ϯ␴ the superconductor, where hϭ0, ⌬ϭ⌬ , and in the ferro- ˆ 0 ˆ 3 ͑ * ˆ Ϫ ˆ ϩ ; ˆ Ϯ ˆ 1 ˆ 2 ; 0 magnet, where ⌬ϭ0, hϭh(x), matching the solutions at the ˇ ϭ␴  ␴  ␴ ϩ ␴  ␴ ͒ ͑ ͒ hЌ ˆ 0 ͑hx ˆ 1 ˆ 0 hy ˆ 2 ˆ 0 ; 10 F/S interface. In order to find the solution in the ferromagnet one needs ˇ ϭ ␴  ␴  ␴ hz hz ˆ 0 ˆ 3 ˆ 3 . to know the detailed character of the magnetic inhomogene- Low Temp. Phys. 27 (99–10), September–October 2001 Kadigrobov et al. 763

␭ Ӷ ity. A quantitative theory can be formulated only in case the 2; this allows us to develop a perturbation theory in D 1 ␭ Ӷ magnetic structure is known in the experiment of interest. In and R 1. Therefore, while solving this equation inside the the absence of any precise information about the magnetic ferromagnet we assume the probability of the spin-flip scat- structure of the samples used in existing experiments, we tering to be small. turn to illustrative examples of magnetic disorder and restrict For both cases, rather simple but cumbersome ourselves to making only qualitative comparisons with ex- perturbation-theory calculations show that far inside the fer- periments. We will consider two such examples. romagnet, where the magnetization is already homogeneous 1. The spin-splitting magnetic scattering is due to a mul- but one is still in the ballistic region ͑at the distance b in Fig. tidomain structure with the magnetizations in the magnetic 2͒, the retarded and advanced parts of the Green’s function ͑ ͒ (␣) ␣ϭ domain near the F/S interface LD region in Fig. 2 and in the gˆ , (R,A) can be written as follows: rest of the ferromagnet being collinear but of opposite direc- ͑␣͒͑ ͒ϭ␶ ␣͑ ͒ϩ␶ ␣͑ ͒ϩ͑␶ Ϫ ˆ ͒ gˆ a b ˆ a˜gs b ˆ s˜ga b ˆ s 1 sign px tions; the width LDW of the domain wall between them is small compared with the ballistic magnetic length L ͑␣͒͑ ͒ϭ␶ ␣͑ ͒ϩ␶ ␣͑ ͒ϩ␶ ͑ ͒ h gˆ s b ˆ s˜gs b ˆ a˜ga b ˆ asign px 16 ϭបv /I . In this case the spin-flip takes place inside the F 0 (␣)ϭ (␣) ϯ (␣) Ϫ domain wall with a probability amplitude proportional to the where gˆ a,s gˆ (px) gˆ ( px) are the antisymmetric and ␭ ϭ symmetric parts of the Green’s functions: parameter D LDW /Lh . ͑a,s͒ ͑a,s͒ 2. The magnetizations in the magnetic domain near the g↑↑ 0 g↑↓ 0 F/S interface and in the rest of the ferromagnet are noncol- 0 ¯g͑a,s͒ 0 ¯g͑a,s͒ linear, with the domain wall between them of a negligible ␣ ↓↓ ↑↓ ˜g ϭ , width. In this case the spin flip takes place due to Rabi os- a,s ͩ ͑a,s͒ ͑a,s͒ ͪ g↓↑ 0 g↓↓ 0 cillations in the L region, the probability amplitude of that ͑ ͒ ͑ ͒ D 0 ¯g a,s 0 ¯g a,s being proportional to the noncollinearity of the magnetiza- ↑↓ ↑↓ ␭ ϭͱ 2ϩ 2 (a,s)ϭ Ϯ Ϫ ␧ tions in the neighboring domains, R hx hy/I0 . where g␴,␴Ј 1/2(g␴␴Ј(px) g␴␴Ј( px)); for the case ͑ ͒ Ӷ͉⌬͉ ␶ ␶ Inserting dimensionless variables into Eq. 13 , one sees the antisymmetric ˆ a and symmetric ˆ s matrices that solutions of this equation in the ferromagnet region are ͑which match the normal and anomalous components of the ␭ ␭ ␣ controlled by the parameter D for case 1 and by R for case Green’s function gˆ (b;px) are as follows:

␸͒ ␸͒ 0 ta exp͑i 0 ra exp͑i Ϫ ␸͒ Ϫ Ϫ ␸͒ ta exp͑ i 0 ra exp͑ i 0 ␶ˆ ϭi sign p ͑17͒ a xͩ 0 Ϫr* exp͑i␸͒ 0 Ϫt exp͑Ϫi␸͒ ͪ a a ͑Ϫ ␸͒ Ϫ ͑Ϫ ␸͒ ra* exp i 0 ta exp i 0 and ␸͒ ␸͒ 1 ts exp͑i 0 rs exp͑i Ϫ Ϫ ␸͒ Ϫ ␸͒ ts exp͑ i 1 rs exp͑ i 0 ␶ˆ ϭ . ͑18͒ s ͩ 0 Ϫr* exp͑i␸͒ 1 t exp͑i␸͒ ͪ s s Ϫ ͑Ϫ ␸͒ Ϫ ͑Ϫ ␸͒ rs* exp i 0 ts exp i 1

Here t ϭit ϩt and r ϭir ϩr are the probability ampli- 0 sf a s sf a s hϩ* 2LD tudes for an electron incident on the magnetically inhomoge- r ϭϪ sin ͑20͒ s l ͉n ͉L neous region to be reflected back as a hole with the same and 0 x h with the opposite direction of its spin, respectively ͉͑t ͉2 sf ͑Eqs. ͑19͒ and ͑20͒ are written for 1у͉n ͉ӷ␲L /L ͒. ϩ͉r ͉2ϭ1; see Fig. 1͒. x DW h sf Equation ͑16͒ together with Eqs. ͑19͒, ͑20͒ show that For the domains with collinear magnetization ͑case 1͒ spin-flip scattering at the magnetization inhomogeneity ac- one has companied by Andreev reflection at the F/S interface pro- ͑ ͑ ͒ ␲ 2 ␲ duces a new triplet order parameter Eq. 1 and a new nor- 1 LDW LDW Ϸ Ϫ ͩ ͪ Ϸ Ϸ ϭ ͑ ͒ mal singlet correlation function G␴ Ϫ␴ proportional to the ta 1 , ra 0;ts 0,rs ͉ ͉ . 19 , 2 Lhnx Lh nx spin-flip probability amplitude rsf . In contrast to the conven- tional singlet order parameter, this triplet order parameter ͑ ͒ For the domains with noncollinear magnetization case 2 does not decay exponentially at distances from the F/S inter- one has face greater than the magnetic length L in the diffusive I0 ͒ ͑0͒ ferromagnet. This fact can be proved for two cases: 1 l0 2LD hz 2LD Ӷ ͒ ӷ ͒ t ϭϪcos , r Ӎ0; t ϭ sin , Lh , and 2 l0 Lh . In case 1 one neglects the terms qua- a ͉n ͉L a s l ͉n ͉L ͉ ͉Ӷ ͑ ͒͑ x h 0 x h dratically small in rsf 1 in the Usadel equation 12 that 764 Low Temp. Phys. 27 (99–10), September–October 2001 Kadigrobov et al.

͑ ͒ is, if one neglects the terms quadratic in F␴,␴ and G␴,Ϫ␴ and d dGˇ tr ͒ ͩ ˇ ͑tr͒ ͪ ϩ͓␧␶ ˇ ͑tr͔͒ϭ ͑ ͒ their derivatives . In this case, for the ‘‘usual’’ components D G ˇ 3 ,G 0. 24 ␣ ␣ dx dx F␴,Ϫ␴ and G␴,␴Ј one gets the conventional Usadel equation ␣ From the above considerations it follows that Eq. ͑22͒ that shows exponential decay of F␴,Ϫ␴ at distances greater ͑ ͒ ␣ ͑ ͒ Ӷ than LI Ref. 2 . As for the F␴ ␴ components, the equations and Eq. 24 are valid for the both cases l0 Lh and 0 , ӷ for them show a slow variation of these components at such l0 Lh . ͑ ͒ distances because these equations have no terms proportional We obtain the boundary conditions for Eq. 24 at dis- ӷ to I . The latter is a mathematical manifestation of the fact tances from the F/S interface of the order of l0(l0 LD) for 0 ប ␧ӷ that these correlation functions are associated with such a the case vF / l0 that permits us to neglect the term pro- portional to ␧ in the Eilenberger equation ͑22͒ and rewrite it scattering process under which the incident electron is trans- 20 formed into a hole without changing the direction of the spin, as the following equation: ץ and hence this electron-hole transformation requires no v t n ˇgtrϭgˇ tr͗gˇ tr͘Ϫgˇ tr͗gˇ tr͘, ͑25͒ R a a s s aץ change in the magnetic energy (I0) of the quasiparticles, as F 0 is qualitatively explained in the Introduction. ͒ ӷ ӷ where In case 2 (l0 Lh) at distances x l0 the Green’s func- tions behave in the same way. In order to see this we start 1 Ӷ tr ϭ ͑ tr͑ ͒Ϯ tr͑Ϫ ͒͒ with solving the Eilenberger equation at distances LD x gˇ s͑a͒ gˇ px gˇ px . Ӷ 2 l0 , where the equation is linear with constant coefficients. ␣ ͑ ͒ It is straightforward to see that the Green functions F␴,Ϫ␴ Averaging Eq. 25 over the momentum direction, one sees ϰ ␣ ϰ ϭ͗ tr͘ 20 exp(ix/(Lhnx)), G␴,Ϫ␴ exp(ix/(Lhnx)), and hence they are that C nxgˇ does not depend on x. Using this fact, Eqs. trϭ ˇ ͑trٌ͒ ˇ ͑tr͒ ͒ ͑ ͒ ͑ ͒ ͑ rapidly oscillating functions of the momentum direction nx at 16 , 18 , and 21 , and the relation gˇ l0G G which ӷ distances x Lh . This means that the averaging with respect couples the isotropic Usadel functions and the anisotropic to the momentum direction results in their decay propor- one at distances greater than l0 , one gets the desired effec- Ӷ 2 tional to Lh /x 1. As to the Green’s functions with the same tive boundary conditions for the Usadel equation ͑24͒ at x ͑ ␣ ␣ ͒ direction of the spin variables G␴,␴ and F␴,␴ , they are Ϸ0 ͑that is in the vicinity of the magnetically inhomoge- slowly varying functions of the momentum direction, and neous region adjacent to the F/S interface͒ as follows: they survive at such distances. Neglecting the rapidly oscil- (␣) d ␣ vF lating components of the Green’s function ͗gˆ ͘ averaged F͑ ,tr͒ͯ ϭ␴ ei␸͉͗n ͉rs͘. ͑26͒ dx ␴,␴ D x over the momentum direction ͑which are small in the param- xϭ0 Ӷ ͒ eter Lh /x 1 , one finds that only the triplet components are ͑ ͒ The exchange energy I0 does not appear in Eqs. 22 , nonzero: ͑24͒ as these equations contain only triplet normal and ͑␣͒ ͑␣͒ anomalous Green’s functions: formally this set of equations G 00F ↑↑ ↓↓ is the same as for a nonmagnetic conductor-superconductor ͑␣͒ ͑␣͒ 0 G¯ ↓↓ ¯F↓↓ 0 diffusive structure if one changes the triplet anomalous ͑␣,tr͒ ϵ ˆ ͑␣,tr͒ϭ ͑ ͒ ͗gˆ ͘ G ͑␣͒ ͑␣͒ . 21 Green’s functions to the singlet ones and uses the boundary ͩ 0 F G 0 ͪ ↓↓ ↓↓ conditions ͑26͒. From here it obviously follows that the spin- ͑␣͒ ͑␣͒ ¯F↑↑ 00G¯ ↑↑ flip scattering due to magnetic inhomogeneity accompanied by Andreev reflection produces a new ͑triplet͒ order param- Using Eq. ͑21͒, one sees that nonlinear Eilenberger eter ͑1͒͑see also Eq. ͑7͒͒ that decays at distances from F/S equation ͑8͒ splits into two sets of equations for the slowly interface of the same order of magnitude as in nonmagnetic varying components of the Green’s function and the rapidly metal-superconductor structures, L␧ϭͱបD/␧, producing a Ӷ oscillating ones in the region Lh /x 1. The latter are not of long-range proximity effect in the ferromagnet. In the next у interest to us as at distances x l0 they decay exponentially; Section we solve the Usadel equation ͑24͒ using boundary the matrix Eilenberger equation for the slowly varying com- conditions ͑26͒ and find the conductance of such a structure. ponents is as follows: i 4. SOLUTION OF THE EFFECTIVE USADEL EQUATION AND ץ trϩͫ␧␶ ϩ ͗ tr͘ trͬϭ ͑ ͒ THE CONDUCTANCE OF THE STRUCTURE g ˇ 3 gˇ ,gˇ 0, 22ˇ ץ ivFn R t0 The current flowing through the structure under consid- where the reduced matrix Green function ͑see Eq. ͑7͒͒ is eration can be written as follows ͑see, e.g., Refs. 22, and 23͒ ͑␣͒ ͑␣͒ ␴ dGˆ ͑K,tr͒ dGˆ ͑A,tr͒ g↑↑ 00f ↓↓ ϭ N ͵ ␴ ͩ ˆ ͑R,tr͒ ϩ ˆ ͑K,tr͒ ͪ ␧ j Tr z G G d ͑␣͒ ͑␣͒ 2e C dx dx 0 ¯g↓↓ ¯f ↓↓ 0 ͑␣,tr͒ ͑27͒ gˆ ϭ ␣ ␣ . ͑23͒ ͩ 0 f ͑ ͒ g͑ ͒ 0 ͪ ↓↓ ↓↓ ␴ ␴ where N is the conductivity of the normal metal; C is a ͑␣͒ ͑␣͒ 4ϫ4 matrix ␴ˆ  ␴ . ¯f ↑↑ 00¯g↑↑ 0 z According to the relation17,21 In the dirty limit, Eq. ͑22͒ reduces to a set of Usadel Gˇ ͑tr͒Gˇ ͑tr͒ϭ1ˇ ͑28͒ equations for the Green’s functions Gˇ ͑␣,tr͒ ͓see Eq. ͑21͔͒ that reads the Keldysh function Gˆ (K,tr) reads Low Temp. Phys. 27 (99–10), September–October 2001 Kadigrobov et al. 765

Gˆ ͑k,tr͒ϭGˆ ͑R,tr͒qˆ ϪqˆGˆ ͑A,tr͒ ͑29͒ where qˆ is a diagonal matrix

q↑1 000

0 q↓2 00 qˆ ϭͩ ͪ , ͑30͒ 00q↓3 0

000q↑4 while the components of retarded and advanced Green’s functions ͑21͒ satisfy the relations FIG. 3. Temperature dependence of the normalized excess conductance ͑see ͑␥,tr͒ ͑␥,tr͒ϩ ͑␥,tr͒¯ ͑␥,tr͒ϭ ͑ ͒ ͑ ͒͒ G␴,␴ G␴,␴ F␴,␴ F␴,␴ 1; Eqs. 39 and 40 . ͑␥,tr͒ϭϪ¯ ͑␥,tr͒ ͑ ͒ G␴,␴ G␴,␴ 31 The solution of the Usadel equation ͑35͒ with the bound- ␥ϭ ˆ (R,tr) ( R,A). Matrix qˆ together with G satisfies matrix ary condition given by Eq. ͑36͒ at xϭ0 and by ⌰␴ϭ0atx equation ͑24͒. ϭL,is Following the reasoning of Ref. 1 and using Eqs. ͑27͒– v ͉͉͗n ͉r ͉͘ sinh͓k͑␧͒͑xϪL͔͒ ͑30͒, one can rewrite Eq. ͑27͒ for the current as follows: ⌰ ϭ␴ F x s ͑ ͒ ␴ ͑␧͒ ͑ ͑␧͒ ͒ , 38 ␴ Dk cosh k L N 1 jϭ ͵ d␧͚ ͑ f ␴͑L͒Ϫ f ␴͑0͒͒ , ͑32͒ ␧ ϭ ϩ ͱ␧ ប 32 ␴ m␴͑L͒ k͑ ͒ ͑1 i͒ / D. ͑ ͒ where f ␴(x) is the distribution function for electrons with Equation 38 shows that the superconducting correla- spin ␴ϭ(↑,↓), and tions due to the spin-splitting processes in the magnetic in- homogeneous region decay exponentially in the ferromagnet L dx ϭͱប ͑ ͑␧͒ϭ ͵ ͑ ͒ and vanish at distances of order LT D/kBT for energies m␴ Ϫ ͑R,tr͒ ͑A,tr͒Ϫ ͑R,tr͒ ͑A,tr͒ . 33 ␧ϳ ͒ 0 1 G␴ G␴ F␴ F␴ kBT , corresponding to the superconducting correlation length in nonmagnetic materials. ˆ (R,tr͒ Therefore, the current is determined by G , the triplet Inserting Eq. ͑38͒ into Eq. ͑37͒, one obtains an excess R ¯ R Green’s functions of which for normal pairing G␴,␴ , G␴,␴ conductance that can be expressed as and anomalous pairing FR ,¯FR can be parameterized ͑see ␴,␴ ␴,␴ ␦G/G ϭ␥ f ͑T/T ͒, ͑39͒ Eq. ͑31͒͒ in the standard way as follows: 0 Th where cosh͑⌰␴͒ sinh͑⌰␴͒exp͑i␹␴͒ ˆ R ϭͩ ͪ ␥ϭ͉ ͉ ͉ ͉2 ͒2 G␴,␴ , ͗ nx rs͘ ͑L/l0 Ϫsinh͑⌰␴͒exp͑Ϫi␹␴͒ Ϫcosh͑⌰␴͒ ͑ ͒ 34 and f (T/TTh) is a dimensionless function, the temperature dependence of which is presented in Fig. 3: where ⌰␴ and ␹␴ are complex functions; the function ␹␴ does not contribute to the conductance ͑see Eq. ͑37͒ below͒. 1 ϱ Ϫ2 2 Using the parameterization ͑34͒ and Eqs. ͑24͒ and ͑26͒, f ͑x͒ϭ ͵ dt cosh ͑t /2x͒ x 0 one gets both the Usadel equation and its boundary condi- tions as sinh͓2͑1ϩi͒t͔Ϫ2͑1ϩi͒t ϫͩ Re 2 ͑ Ϫ ͒ 2 2͓͑ ϩ ͒ ͔ d ⌰␴ 4 i 1 t cosh 1 i t ប Ϫ ␧⌰ ϭ ͑ ͒ D 2 2i ␴ 0, 35 dx sinh 2tϪsin 2t ϩ ͪ ͑ ͒ 2͉ ͑ ϩ ͒ ͉2 . 40 d v 4t cosh 1 i t ⌰ ͯ ϭ␴ F ͉͉͗ ͉ ͉͘ ͑ ͒ ␴ n r . 36 dx D x s Using experimental values of the parameters taken from xϭ0 Ref. 10, Dϭ100 nm2/s and T/T ϭ50, and making the rea- ͓ Th Here rs is the magnetic spin-flip scattering amplitude see sonable assumption that r ϳ0.1, we find that our result for ͑ ͒ ͑ ͔͒ ͑ ͒ ͑ ͒ sf Eq. 19 , 20 . Equations 35 and 36 are linear due to the the excess resistance, ␦RϷϪ10 ⍀, is in agreement with the smallness of the amplitude for magnetic spin-flip scattering, experiment. The temperature dependence of the excess con- ͉ ͉Ӷ Ӷ⌬ rs 1, and are valid in the temperature interval kBT , ductance in the range TϳT is shown in Fig. 3. For higher ϭប 2Ӷ⌬ Th which includes the Thouless energy kBTTh D/L . ϳ⌬ ӷ ϭប 2 temperatures, T /kB TTh D/(kBL ), our theory is not In order to calculate the conductance we follow Ref. 1 ⌬ϳ valid, and contributions of order kBT/ 1 can modify the and find that the excess conductance can be written as temperature dependence of the resistance. Additional mea- 1 ͑ -f 1 L surements around the Thouless temperature where the prox ץ G 1 ϱ␦ 0 2 ϭϪ ͚ ͵ d␧ ͫ ͵ dx͑Re ⌰␴͒ ͬ, imity effect is most pronounced͒ would permit a comparison ␧ Lץ G 16T ␴ϭϪ Ϫϱ 0 1 0 with the temperature dependence coming from the long- ͑ ͒ 37 range proximity effect described by our theory. However, where G0 is the conductance of the ferromagnetic part of the additional investigations of the magnetic structure of the F/S ␧ structure; f 0( ) is the Fermi distribution. interface are needed to carry out a complete comparison with 766 Low Temp. Phys. 27 (99–10), September–October 2001 Kadigrobov et al. the theory. Multidomain ferromagnets suitable for these stud- 3 K. Kawaguchi and M. Sohma, Phys. Rev. B 46, 14722 ͑1992͒. ies can be created in various ways. It was recently 4 J. S. Jiang, D. Davidovic, D. H. Reich, and C. L. Chien, Phys. Rev. Lett. 24 74,314͑1995͒. demonstrated that grain boundaries, magnetic inhomogene- 5 ities ͑including domains with nonparallel magnetization͒ can R. J. Soulen Jr., J. M. Byers, M. S. Osofsky, B. Nadgorny, T. Ambrose, S. F. Cheng, P. R. Broussard, C. T. Tanaka, J. Nowak, J. S. Moodera, and J. be introduced in a predetermined position in a ferromagnet M. D. Coey, Science 282,85͑1998͒. film by controlling the epitaxial growth. Experiments in 6 A. M. Goldman, V. Vas’ko, P. Kraus, and K. Nikolaev, J. Magn. Magn. which such magnetic inhomogeneities are intentionally cre- Mater. 200,69͑1999͒. ated would permit long-range proximity effects to be studied 7 L. Lazar, K. Westerholt, H. Zabel, L. R. Tagirov, Yu. V. Goryunov, N. N. ͑ ͒ in well-characterized ferromagnet—superconductor struc- Garif’yanov, and I. A. Garifulin, Phys. Rev. B 61,3711 2000 . 8 V. T. Petrashov, I. A. Sosnin, I. Cox, A. Parson, and C. Troadec, Phys. Rev. tures. Lett. 83, 3281 ͑1999͒. In conclusion, we have shown that spin-splitting scatter- 9 V. T. Petrashov, I. A. Sosnin, I. Cox, A. Parson, and C. Troadec, J. Low ing related to magnetic inhomogeneities modifies the spin- Temp. Phys. 118, 689 ͑2000͒. selection rule governing Andreev reflections at a ferromag- 10 M. D. Lawrence and N. Giordano, J. Phys.: Condens. Matter 8, L563 netic normal metal-superconductor interface. As a result a ͑1996͒. 11 M. Giroud, H. Courtois, K. Hasselbach, D. Mailly, and B. Pannetier, Phys. long-range proximity effect, due to correlations between ͑ ͒ ͑ Rev. B 58, R11872 1998 . spin-aligned electrons and holes, appears a spin-triplet prox- 12 F. Zhou and B. Spivak, cond-mat/9906177. imity effect͒. Estimations of the value of the excess conduc- 13 F. S. Bergeret, K. B. Efetov, and A. I. Larkin, Phys. Rev. B 62, 11872 tance are consistent with experiments.8–11 For applications, ͑2000͒. an important consequence of this phenomenon is that the 14 J.-X. Zhu and C. S. Ting, Phys. Rev. B 61, 1456 ͑2000͒. 15 proximity effect can be stimulated by orders of magnitude by A. Kadigrobov, R. I. Shekhter, and M. Jonson, Europhys. Lett. 54,394 ͑ ͒ intentionally produced magnetic inhomogeneity in the 2001 ; cond.-mat/00124347. 16 F. S. Bergeret, A. F. Volkov, and K. B. Efetov, Phys. Rev. Lett. 86, 4096 sample. ͑2001͒; cond-mat/0011425. 17 G. M. Eilenberger, Z. Phys. 214, 195 ͑1968͒. We acknowledge useful discussions with E. V. Bezuglyi 18 K. D. Usadel, Phys. Rev. Lett. 25, 507 ͑1970͒. and Z. G. Ivanov. We are grateful to Z. G. Ivanov for calling 19 J. Rammer and H. Smith, Rev. Mod. Phys. 58, 323 ͑1986͒. to our attention experimental possibilities for observing the 20 M. Yu. Kuprianov and V. F. Lukichev, Zh. E´ ksp. Teor. Fiz. 94,139͑1988͒ effect predicted by our theory. ͓Sov. Phys. JETP 67, 1163 ͑1988͔͒. 21 A. I. Larkin and Yu. N. Ovchinnikov, Zh. E´ ksp. Teor. Fiz. 68,1915͑1975͒; ͑ ͓͒ ͑ ͒ ͑ ͔͒ * 73,299 1977 Sov. Phys. JETP 41,960 1975 ; ibid. 46,155 1977 . E-mail: [email protected] 22 ͒ W. Belzig, F. K. Wilhelm, C. Bruder, G. Scho¨n, and A. D. Zaikin, Super- 1 A short formulation of this prediction was published as a Letter in Ref. 15. lattices Microstruct. 25, 1252 ͑1999͒. Analogous prediction based on a somewhat different approach was simul- 23 taneously made in Ref. 16. E. V. Bezuglyi, E. N. Bratus’ and V. P. Galaiko, Fiz. Nizk. Temp. 25,230 ͑1999͓͒Low Temp. Phys. 25,167͑1999͔͒. 24 R. Mathieu, P. Svedlindh, R. A. Chakalov, and Z. G. Ivanov, Phys. Rev. B 62, 3333 ͑2000͒. 1 C. J. Lambert and R. Raimondi, J. Phys.: Condens. Matter 10, 901 ͑1998͒. 2 E. A. Demler, G. B. Arnold, and M. R. Beasly, Phys. Rev. B 55, 15174 This article was published in English in the original Russian journal. Repro- ͑1997͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Formation of d-wave superconducting order in a randomly doped lattice V. M. Loktev* and Yu. G. Pogorelov

Bogolyubov Institute for Theoretical Physics of the National Academy of Science of Ukraine, 14b, Metrologichna Str., Kiev, 03143, Ukraine; CFP and Departamento de Fı´sica da Faculdade de Cieˆncias da Universidade do Porto, Rua do Campo Alegre, 687, 4169-007 Porto, Portugal ͑Submitted May 14, 2001͒ Fiz. Nizk. Temp. 27, 1039–1052 ͑September–October 2001͒ We consider the interplay between superconducting coupling and dopant impurity scattering of charge carriers in planar square lattice systems and examine the physical conditions ͑doping level, temperature, local symmetry of coupling and scattering potentials͒ necessary in this model system to obtain a d-wave superconducting order, like that observed in real doped cuprate HTSC materials. Using the Lifshitz model for the disorder introduced into the system by dopants, we analyze also the nonuniform structure of such a d-wave parameter, including both its magnitude and phase variation. The results indicate that d-wave superconductivity becomes possible in a doped metal until it is destroyed at too high doping levels. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1401186͔

1. INTRODUCTION did not include one of the most essential features of HTSC systems: the fact that they cannot be non-impure. Studies of the effect of impurities and defects on super- In other words, most of theoretical approaches to HTSC ͑ ͒ ͑ ͒ conducting SC properties of metals including SC alloys are based on the concept formulated and applied in the afore- began practically as early as the BCS theory had been con- mentioned pioneering works:1,2 one starts from an ideal ͑2D 1 structed. In particular, the classical papers by Anderson and or quasi-2D͒ metal with a given Fermi energy ␧ , defined by 2 F by Abrikosov and Gor’kov indicated a substantial difference the density of free carriers, and then considers the perturba- between magnetic and nonmagnetic impurities in supercon- tion of an independently existing SC condensate by some ductors. While the addition of nonmagnetic impurities has extrinsic ͑magnetic or nonmagnetic͒ impurities. Their action, practically no effect on the value of the transition critical local or global, affects the preformed and condensed singlet temperature T , the presence of spin on the impurity atom c pairs. Of course, this formulation is reasonable but it lacks an ͑leading to the Kondo effect in a normal metal͒ results in essential element for the conductance in copper oxides: al- pair-breaking, that is, it transforms a singlet Cooper pair into most all HTSCs are doped metals, where ͑like the doped an unstable triplet and rapidly suppresses T . All of the c semiconductors͒ each carrier is provided by insertion of a known studies have considered the s-type, or isotropic, SC donor or acceptor into the system. In turn, this implies that order ͑apart from heavy-fermion systems, where it is the HTSCs are intrinsically impure systems with an inherent p-type order and T is extremely low͒ and, correspondingly, c disorder,1͒ and the number of impurity ͑‘‘foreign’’ included͒ an isotropic gap near the Fermi level. The discovery of high-T superconductivity ͑HTSC͒ in atoms can not be less than the number of charge carriers, at c ӷ copper oxides confronted physicists with a number of prob- least. While in normal metals one has the condition kFl 1 ͑ lems which still remain a challenge for the theory. Undoubt- kF is the Fermi momentum, and l is the carrier mean free ͒ 18 edly, this includes the issue of the HTSC mechanism, the path between collisions with impurity atoms , it turns to ϳ strong dependences of many ͑both SC and normal͒ proper- kFl 1 in HTSCs, and they belong to the family of ‘‘bad’’ ties of copper oxides on the charge-carrier concentration, the metals with both kF and l determined by the doping. specifics of weakly doped systems ͑first of all, the existence Perhaps the first attempt to consider in a self-consistent ͒ way the characteristic tendencies for HTSC, pairing of the of a pseudogap at temperatures above Tc , the formation of stripe structures, etc. ͑see, e.g., the review articles͒.3–6 Such carriers and their localization on impurity atoms, was made 19 a problem is also presented by the impurity effect on the SC in the authors’ previous paper. It discussed the phase dia- properties of HTSC systems. These differ from the ‘‘old’’ or gram of doped 2D metal in presence of s-wave pairing and ‘‘conventional’’ superconductors not only in their higher Tc showed that SC is possible neither at low impurity concen- Ͻ ϳ␧ ͑ and the d-wave anisotropy of their order parameter but also trations c c0 0 /W when all the carriers are localized ␧ in the fact that here magnetic and nonmagnetic impurities near impurities with localization energies 0 much less than ͒ change their roles: the former are weak suppressors of Tc the bandwidth W, so that c0 is typically few percent nor at ͑Refs. 7 and 8͒, while the latter ͑in particular, Zn substituting too high c ͑when the pair inverse lifetime times ប exceeds for Cu in cuprate layers͒ lead to a fast decay of HTSC.7,10,11 the SC gap͒. There, in general, the self-consistency is related Many aspects of impurity effects in superconductors with either to the SC order parameter ͑like the common Bardeen- anisotropic ͑including d-wave͒ pairing have already been Cooper-Schrieffer or Bogolyubov-de Gennes treatments͒ and studied theoretically in Refs. 12–17, using a range of models to the chemical potential. and approximations. However, these ͑and many other͒ papers The present work is aimed at extending the approach of

1063-777X/2001/27(9–10)/10/$20.00767 © 2001 American Institute of Physics 768 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Loktev and Yu. G. Pogorelov

␧ Ͻ␮ ͑ Ref. 19 to the case of d-wave SC coupling and to trace the tion D to hold and a BCS shell to exist the alternative ␮Ͻ␧ formation of the corresponding order parameter. It is moti- D , possible for underdoped HTSC systems, will be ͒ vated, not least, by an apparent controversy between the ex- considered elsewhere . The impurity perturbation VL ex- perimental evidence for d-symmetry of the order parameter presses the shift of the on-site electronic energy at a random in HTSCs20–22 and the theoretical claim that anisotropic pair- dopant site p, where the negative sign takes explicit account ing should not survive in the presence of chaotically distrib- of the carrier attraction to the ionized dopant, and, for sim- uted isotropic scatterers.23 For the sake of simplicity, we re- plicity, we consider this perturbation localized on a single Ͼ ϭ strict consideration to the doping range c c0 , where the site. With the usual BCS ansatz: cÀk,↓ck↑ ͗cÀk,↓ck,↑͘ ϩ␸ self-consistency is only relevant for the SC order parameter, k , and in neglect of terms quadratic in the pair fluctua- 2͒ ␮ ␸ ͑ ͒ ϭ Ϫ␮ while the chemical potential can be put in the form tions k , Eq. 1 leads to a bilinear form HЈ H N: Ϸ␧ Ϸ F 3cW/4. Then we distinguish between two types of impurity effects by doping. The first, the so-called homoge- ϭ ␰ ϩ Ϫ ⌬ ϩ ͒ HЈ ͚ ͚ͫ kc ␴ck,␴ ͑ kcϪk,↓ck,↑ h.c. k, neous effects, are displayed by translationally invariant k ␴ single-particle Green functions ͑SPGFs͒. They have been 24 V studied earlier by various means but, as a rule, by introduc- Ϫ L Ϫi͑kЈϪk͒ p ϩ ͑ ͒ ͚ e • c ␴ck,␴ͬ. 2 kЈ, ing the disorder through a single parameter VA of Anderson’s N p,kЈ,␴ model.24 In contrast, we employ Lifshitz’s model of disorder,25 characterized by two independent parameters: c Here the gap function is defined by the self-consistency re- and the impurity potential VL . They produce an equivalent lation ϳͱ Ϫ VA c(1 c)VL , but not vice versa. Within this model, V␥ more adequate for doped HTSC systems, we conclude the ⌬ ϭ k ␥ ͑ ͒ k ͚ kЈ͗ckЈ,↑cϪkЈ,↓͘, 3 persistence of the d-wave order parameter under homoge- N kЈ neous impurity effects. Also we explicitly consider the other type of effects, inhomogeneous, due to local variations of the extending the common BCS gap equation to the d-wave order parameter near impurity centers. This involves two- case. A nonuniform system can be treated within the formal- 14,19 particle Green functions ͑TPGFs͒ besides the usual SPGFs, ism used formerly for impurity problems in SC by pass- ␺ϩϭ ϩ ␺ and yields a possible limitation on SC at high enough dopant ing to the Nambu spinors k (ck,↑ ,cÀk,↓) and k and de- concentrations. fining the Fourier-transformed matrix Green function ͑GF͒ At least, we would like to acknowledge the great honor 0 and pleasant opportunity for us to publish this paper in the ˆ ͑␧͒ϵ͗͗⌿ ͉⌿ϩ ͘͘ ϭ ͵ i͑␧Ϫi0͒t͕͗⌿ ͑ ͒ ⌿ϩ ͖͘ Gk,kЈ k kЈ ␧ e k t , kЈ dt. Low Temperature Physics issue dedicated to the memory of Ϫϱ the outstanding physicist L. V. Shubnikov, whose contribu- ͑4͒ tion to low-temperature physics in general and to supercon- ˆ ϫ ͗ ͘ ductivity in particular cannot be overestimated. Here A denotes a 2 2 matrix in Nambu indices, ... is the quantum statistical average, and ͕a(t),b(0)͖ is the anticom- mutator of Heisenberg operators. In the GFs below we omit 2. HAMILTONIAN AND GREEN FUNCTIONS their explicit dependence on energy ␧ but distinguish be- tween their diagonal and nondiagonal forms in the Nambu We start from the model electronic Hamiltonian in the ͑N͒ and momentum ͑M͒ indices. Then, applying the Heisen- ͒ ͑ ϭ ␺ ץ ␺ץband representation, ប berg equation of motion i k / t ͓HЈ, k͔ in Eq. 4 ,we arrive at an equation of motion of the Dyson type for the ϭ ␧ ϩ H ͚ ͭ ͚ kc ␴ck,␴ k, SPGFs: k ␴

1 ˆ ϭ ˆ ͑0͒␦ Ϫ ˆ ͑0͒ ˆ i͑kϪkЉ͒ p ˆ ͑ ͒ ϩ ϩ Gk,kЈ Gk k,kЈ Gk V ͚ e • GkЉ,kЈ 5 Ϫ ͚ ͫ V␥ ␥ c n cϪ ↓c ↑ k kЈ k,↑ Ϫk,↓ kЈ, kЈ, p,kЉ N kЈ ˆ (0)ϭ ␧Ϫ␰ ␶ Ϫ⌬ ␶ where the unperturbed SPGF Gk ( kˆ 3 kˆ 1 i͑kЈϪk͒ p ϩ ϪV ͚ e • c c ␴ͬ , ͑1͒ Ϫ1 L ␴ k, ͮ ϩ ˆ ϭ ␶ ␴ kЈ, i0) , and the scattering matrix V VL ˆ 3 includes the p, ␶ Pauli matrices ˆ i . ϩ ͑ ͒ where ck,␴ and ck,␴ the Fermi operators for a charge carrier For a disordered system, the relevant observable char- with wave vector k and spin ␴. The simplest band energy acteristics are described by the so-called self-averaging GFs, ␧ ϭ Ϫ ϩ ϭ k 4t 2t(cos akx cos aky), with full bandwidth W 8t,is whose values for all particular realizations of disorder turn expressed through the amplitude t of carrier hopping be- practically nonrandom, equal to those averaged over tween nearest-neighbor sites.3͒ ͑of the total N in the lattice, disorder.26 The most important example of such a function is ͒ ˆ ϵ ˆ with a lattice constant a . The parameter V models the attrac- the M-diagonal SPGF, Gk Gk,k . The general solution for tion between two carriers with opposite spins on such sites, Eq. ͑5͒ in this case can be written ͑see Appendix E͒ as ␥ ϭ Ϫ ␪ ␧2 Ϫ␰2 the factor k (cos akx cos aky) ( D k) has d-wave sym- ␰ ˆ ϭ͕͓ ˆ ͑0͔͒Ϫ1Ϫ⌺ˆ ͖Ϫ1 ͑ ͒ metry and is effective only for quasiparticle energies k Gk Gk k , 6 ϭ␧ Ϫ␧ ␧ k F smaller than the ‘‘Debye energy’’ D . The latter is ⌺ˆ understood as a characteristic energy of the intermediate where the self-energy matrix k is given by the so-called ͑Fro¨lich͒ boson, and in what follows we suppose the condi- fully renormalized group expansion ͑GE͒14,19,27,28 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Loktev and Yu. G. Pogorelov 769

␳ Ϸ ␲ Ϫ Ϫ where 0 4/( W) is the normal Fermi density of states of a ⌺ˆ ϭϪcVˆ ͓1ϩGˆ Vˆ ͔ 1ϫͭ 1ϩc ͚ ͓Aˆ e ik•n ͑ ͒ ␧ Ͻ k 0n doped quasi-2D metal with W/2. Accordingly, the real nÞ0 F part of Gˆ (0) is ϩAˆ Aˆ ͔͓1ϪAˆ Aˆ ͔Ϫ1ϩ ͮ . ͑7͒ ␲ ⌬ 0n n0 0n n0 W ¯ ReGˆ ͑0͒ϭ␧␳ ͫ Ϫ ␪͑⌬Ϫ␧͒arccosh ͬ. ͑14͒ 0 ␮͑WϪ␮͒ ⌬ ␧ Here the integrated SPGF matrix Gˆ ϭNϪ1⌺ Gˆ , and the k k Then Eqs. ͑13͒ and ͑14͒ can be unified into a single analytic matrices Aˆ of indirect interaction between scatterers at 0n form: sites 0 and n are ␲ ⌬ ␲ ͑ ͒ W Ј Ϫ ˆ 0 ϭ␧␳ ͫ Ϫ ͩ Ϫ ͪͬ ͑ ͒ ˆ ϭϪ ˆ ik n ˆ ͓ ϩ ˆ ˆ ͔ 1 ͑ ͒ G 0 arccosh i , 15 A0n V ͚ e • GkЈ 1 GV . 8 ␮͑WϪ␮͒ ⌬ ␧ 2 kЈÞk since at ␧Ͼ⌬ one has arccos (⌬/␧)ϭi͓␲/2Ϫarcsin(⌬/␧)͔, The restriction to kЈÞk in the single summation in Aˆ 0n thus restoring Eq. ͑13͒. But it is just the growth of the ͑real͒ should be complemented by kЉÞk,kЈ for double summation arccosh term at ␧Ͻ⌬ that permits the existence of a low- in the product Aˆ Aˆ , but such restrictions can be ignored Ϫ 0n n0 energy (␧ Ӷ⌬) resonance feature in Re(1ϩGˆ Vˆ ) 1 and ˆ ˆ ˆ 27 res in A0nA0nAn0 and the higher-degree terms resulting from hence in ␳͑␧͒. Such a resonance was discussed previously14 ͑ ͒ expansion of the right-hand side of Eq. 7 . for a d-wave SC with low enough concentration c ͑so that Many observable characteristics of the SC state follow Gˆ ϷGˆ (0)͒ of ‘‘foreign’’ impurities producing a strong enough from the GFs, in the spectral theorem representation perturbation VL , and it is similar to the known low- ϱ d␧ 1 frequency resonance by heavy impurities in acoustic phonon ͗ ͘ϭ ͵ ͗͗ ͉ ͘͘ ␤ϭ ͑ ͒ ab ␤͑␧Ϫ␮͒ Im b a ␧ , , 9 29 Ϫϱe ϩ1 T spectra. However, in the situation of interest here, when both V and c are not small, Gˆ can be substantially modified where the chemical potential ␮ is defined by the overall elec- L compared to Gˆ (0), and this is expressed in a very compli- tron concentration cated way by Eq. ͑7͒. To simplify the task, certain self- 1 ϱ d␧ consistent procedures, like the CPA method, quite useful in ϭ ͵ ␶ ˆ ͑ ͒ c ͚ ␤͑␧Ϫ␮͒ Im Trˆ 3Gk . 10 30 N k Ϫϱe ϩ1 the theory of normal metals, can be employed. A similar approach was previously proposed for an s-wave SC doped On the other hand, c is just the concentration of dopant cen- system,19 and here we begin with the analysis of a self- ϭ Ϫ1⌺ ters which give rise to carrier scattering: c N p1 and the consistent solution for Eq. ͑5͒ in the d-wave case. carrier concentration only gets close to ͑but never exceeds͒ c in the regime of the doped metal, for c above a certain met- ͑ ␧ allization threshold c0 for a quasi-2D dispersed k it is c0 ϳ Ϫ␲ Ӷ 19 3. UNIFORM DOPING EFFECTS IN THE SELF-CONSISTENT exp( W/4VL) 1. Under this condition, the self- consistency implied by Eq. ͑10͒ is not necessary, and a good APPROACH 4͒ ␮Ϸ ͑ approximation for the chemical potential is 3cW/4 see If the GE series ͑7͒ is restricted to its first term, the ͒ ͑ ͒ Appendix A . Then the gap equation 3 takes the form ⌺ˆ self-energy matrix k in fact becomes independent of k: ␥ ϱ ␧ V k d ⌬ ϭ ␥ ͵ ␶ ˆ ͑ ͒ Ϫ1 k ͚ kЈ ␤͑␧Ϫ␮͒ Im Trˆ 1Gk , 11 ⌺ˆ →⌺ˆ ϭϪ ˆ ͓ ϩ ˆ ˆ ͔ ͑ ͒ Ϫϱ ϩ k cV 1 GV , 16 2N kЈ e 1 ͑ ͒ ͑ ͒ and its solution, discussed in Appendix C for the uniform and substitution of Eq. 16 into Eq. 6 defines the self- ϭ ⌬ ϭ⌬␥ ˆ (sc) case (VL 0), is simply k k , with the ratio r consistent approximation Gk for the M-diagonal SPGF: ϭ2⌬/k T being e1/3 times the s-wave BCS value r B c BCS ˆ ͑sc͒ϭ͕͓ ˆ ͑0͔͒Ϫ1Ϫ⌺ˆ ͑sc͖͒Ϫ1 ͑ ͒ Ϸ3.52. Gk Gk , 17 Another important self-averaging quantity is the inte- ⌺ˆ ͑sc͒ϭϪ ˆ ͓ ϩ ˆ ͑sc͒ ˆ ͔Ϫ1 ͑ ͒ grated SPGF matrix Gˆ itself, since the density of states ␳͑␧͒ cV 1 G V , 18 is just 1 Gˆ ͑sc͒ϭ ͚ Gˆ ͑sc͒ . ͑19͒ 1 N k ␳͑␧͒ϭ ˆ ͑ ͒ k ␲ Im TrG. 12 To solve this system, we first parametrize the self-energy → ˆ → ˆ (0)ϭ Ϫ1⌺ ˆ (0) ͑ ͒ For an unperturbed system VL 0, G G N kGk , matrix 18 : and calculation of the imaginary part of Gˆ (0) within the ⌺ˆ ͑sc͒ϭ⌺ ϩ⌺ ϩ␶ ϩ⌺ ␶ ͑ ͒ nodal point approximation ͑Appendix B͒ leads to the stan- 0 1 ˆ 1 3 ˆ 3 , 20 ⌺ dard d-wave density of states: where the i are some complex-valued functions of energy. ͑ ͒ 1 Then the integration in Eq. 19 within the nodal-point ap- ␳͑␧͒→␳͑0͒͑␧͒ϭ ˆ ͑0͒ ͑ ͒ ␲ Im TrG proximation Appendix D results in ˆ ͑sc͒ϭ ϩ ␶ ϩ ␶ ͑ ͒ ␧␳ ⌬ G G0 G1 ˆ 1 G3 ˆ 3 21 2 0 ϭ arcsinͫminͩ 1, ͪͬ, ͑13͒ ⌬ ␧ with the coefficients 770 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Loktev and Yu. G. Pogorelov

␲ ⌬ϩ⌺ numerically͒ within the whole relevant energy range. It turns ϭ͑␧Ϫ⌺ ͒␳ ϫͫϪ ͩ 1 G0 0 0 ⌬ arccosh ␧Ϫ⌺ especially simple if ͉⌺ ͉ӶcV˜ ͑this proves to hold at least for 2 0 0 ␧Ӷ␧ ͒; then the proper solution to Eq. ͑29͒ is G ⌬Ϫ⌺ res 0 1 W Ϸ⌺ ˜ 2 ͑ ͒ ϩarccosh Ϫi␲ͪ ϩ ͬ, ͑22͒ 0 /cV , and from Eq. 26 we obtain the following equa- ␧Ϫ⌺ ␮͑ Ϫ␮͒ ⌺ 0 W tion for a single important function 0 : G ϭ⌺ ␳ ⌬ ˜␣⌺ 1 1 0 ϭ ␣Ϫ 0 ͑ ͒ ␧Ϫ⌺ sinͩ ␧Ϫ⌺ ͪ 30 2i␲ 0 0 ͫ Ϫ ͱ ␧Ϫ⌺ ͒2Ϫ ⌬ϩ⌺ ͒2ϩͱ ␧Ϫ⌺ ͒2Ϫ ⌬Ϫ⌺ ͒2 ␣ϭ⌬ ͓␲ ˜ 2␳ ͔ӷ ͑ 0 ͑ 1 ͑ 0 ͑ 1 with ˜ / cV 0 1. It defines the self-consistent den- sity of states W ϩ ͬ, ͑23͒ ⌺ ͑␧͒ ␮͑ Ϫ␮͒ ͑ ͒ 1 ͑ ͒ 2Im 0 W ␳ SC ͑␧͒ϭ Im TrGˆ SC ͑␧͒ϭ , ͑31͒ ␲ ␲cV˜ 2 ␮ ͑␧Ϫ⌺ ͒2Ϫ⌬2 Ϫ⌺2 0 /3 1 G ϭ␳ ͫ ln ϩ2⌺ ͬ. ͑24͒ 3 0 WϪ␮ 3 ␧3 at lowest energies. The results of this approach are free of the D infrared logarithmic divergences that appear in the integrals Substituting Eq. ͑21͒ into Eq. ͑18͒, we arrive at of perturbation theory32 and thus allow one to avoid applying heavy field-theoretic methods for a white-noise scattering cV ͓V ͑G ϩG ␶ˆ ͒Ϫ͑1ϩV G ͒␶ˆ ͔ ⌺ˆ ͑sc͒ϭ L L 0 1 1 L 3 3 ͑ ͒ potential,15 whose adequacy to the case of discrete random ͑ ϩ ͒2Ϫ 2͑ 2Ϫ 2͒ . 25 1 VLG3 VL G0 G1 dopants is not clear. Comparing Eqs. ͑25͒, and ͑22͒–͑24͒ with Eq. ͑20͒,weim- The exact value of the density of states at the very center of the gap, ␳(0)ϭ␳(␧→0), is also of a particular interest mediately conclude that ⌺ ϭG ϭ0, or that ⌺ˆ (sc) is in fact 1 1 in view of the known claim about existence of a nonzero N-diagonal, which is extremely important. Physically, this ‘‘universal’’ value ␳(0)ϳc/ln(1/c)ifV is sufficiently means that ͑within the self-consistent approximation linear in L strong.12,17 However, we conclude from Eq. ͑30͒ that in the c͒ the scattering by dopants does not influence the d-wave limit ␧→0: order parameter, and this can be directly related to the fact that the s-symmetry of the impurity perturbation VL is or- ˜␣ ⌺ →␧ͫ1ϩi ͬ, thogonal to the d-symmetry of the SC pairing V. It also 0 ln͑2⌬/˜␣͉␧͉͒ applies to more realistic models of dopant perturbation in HTSC ͑e.g., with plaquette- or dumbbell-like anisotropy͒,31 and hence of the self-consistent density of states at ␧→0: provided that their symmetries do not coincide with that of ͑ ͒ 4˜␣2␳ 0 ͑␧͒ the order parameter. Complications arise when they do coin- ␳͑SC͒͑␧͒→ ͑ ͒ ␲ ͑ ⌬ ␣͉␧͉͒ , 32 cide, as was found for an isotropic perturbation on s-wave In 2 /˜ ⌺ 19 order with all three i being nonzero; hence the apparently vanishes even faster than the unperturbed function ␳(0)(␧)of ‘‘harder’’ d-wave system in fact turns ‘‘easier’’! Eq. ͑13͒. This produces a certain narrow ‘‘quasi-gap’’ ͑not to By using the fact that ⌺ ϭ0 and the relation cosh(x Ͼ 1 be confused with the pseudogap observed at T Tc in the ϩi␲/2)ϭsin x, Eq. ͑22͒ is brought to a very simple form: underdoped regime͒ around the center. For comparison, the ␳͑␧͒ ͑ ͒ ⌬ G ⌬ estimated from the two first terms of Eq. 7 tends to ϭ ␣Ϫ 0 ͑ ͒ ␧ ϳ␧2 ͑ ͒ ␧Ϫ⌺ sinͩ ␲␳ ␧Ϫ⌺ ͪ , 26 zero linearly in with corrections Ref. 14 , while 0 0 0 the field-theoretical analysis15 predicts ␳(␧)→␧␻, with with ␣ϭW⌬/͓␲␮(WϪ␮)͔Ӷ1, while the same comparison the nonuniversal exponent being ͑in our notation͒ ⌺ˆ (sc) ⌺ ⌺ ␻ϭtanh ln ͱ␲2 ⌬W/2cV2 , which is always Ͻ1 and can for the two nonzero components of : 0 and 3 , gives: L even be Ͻ0. ⌺ ͓͑ ϩ ͒2Ϫ 2 2͔ϭ 2 ͑ ͒ 0 1 VLG3 VLG0 cVLG0 , 27 The discrepancy between our results and the aforemen- tioned ‘‘universal’’ behavior originates in the improper use of ⌺ ͓͑ ϩ ͒2Ϫ 2 2͔ϭϪ ͑ ϩ ͒ ͑ ͒ 3 1 VLG3 VLG0 cVL 1 VLG3 . 28 →ϱ ϩ the unitary limit VL in neglect of the 1 VLG3 term in ͑ ͒ ͉⌺ ͉ϳ ͑ ͒ ⌺ ϭϪ From Eq. 28 we estimate 3 cVL ; hence, within the Eq. 27 , leading to the relation 0 c/G0 . But the true ͉␧͉Ӷ␧ ⌺ ϭ ˜ 2 ˜ relevant energy region D the function G3 from Eq. limiting relation is inverse: 0 cV G0 , with V finite for ͑ ͒ ͑ ͒ →ϱ 5͒ 24 is reasonably well approximated by a negative con- VL , and also the unitary limit fails at any finite VL ϭ␳ ␮ Ϫ␮ ͑ ͒ stant g3 0 ln͓ /(W )͔. Then Eq. 27 turns quadratic for when ␧→0. Finally, the existing experimental data do not G0 : confirm the ‘‘universal’’ ␳͑0͒ value but seem to favor the 2 conclusion that there exists a strong low-energy resonance in 11 33 2 c 1 ␳͑␧͒, with a possible quasi-gap at the very center, though G ϩ G Ϫͩ ͪ ϭ0, ͑29͒ 0 ⌺ 0 0 ˜V experimental observations at such low energies of course re- quire extremely low temperatures. ˜ ϭ ϩ Ϸ with V VL /(1 VLg3) VL ln(1/c0)/ln(3c/4c0). The sys- Note, however, that the self-consistent treatment of the ͑ ͒ ͑ ͒ ␧ ϳ⌬ ␲ ␲ tem of equations 26 , 29 fully determines the self-energy low-energy resonance, at res ln(3c/4c0)/( ln 2 ) for ⌺ ͑ ͒ 0 and other uniform physical properties of a disordered the case of self-impurities Fig. 1 , already requires solution d-wave system, and its solution can be found ͑in principle, of the full system of equations ͑26͒, ͑29͒, and, in view of a Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Loktev and Yu. G. Pogorelov 771

FIG. 2. Schematic of consecutive scattering processes ͑see Appendix E͒, bringing an M-nondiagonal TPGE f, through intermediate, partly M-diagonal f Ј and f Ј to fully M-diagonal f Љ: the solid vectors standing for ͑ ͒ 1 2 FIG. 1. Density of states in a d-wave SC system heavy solid line . The thin the operators of the left-hand side get fully nested with the dashed vectors (0) solid line shows the unperturbed ␳ (␧), and the arrow indicates the solu- for the right-hand side. ͓ ϩ 2Ϫ 2 2͔ϭ tion for Re (1 VL G3) VLG0 0, corresponding to the low-energy reso- nance ͑the other solution at ␧Ϸ⌬ is not resonant͒ for the choice of param- ϭ ϭ ϭ eters W 2eV,VL 0.5 eV, c 0.1. Using Eq. ͑33͒, we write 2 V ͑ ϩ Ϫ Ϫ ͒ ␦2ϭ ͚ ͚ ei k1 k2 k3 k4 •n␥ ␥ probable underestimate of this hump ͑like that in normal 3 k1 k2 N n k1 ,k2 ,k3 ,k4 systems͒,28,30 it should be better described by the exact GE ͑ ͒ ϫ͓͗cϪ ↓c ↑cϪ ↓c ↑͘Ϫ͗cϪ ↓c ↑͘ 7 . k1 , k3 , k2 , k4 , k1 , k3 , ˆ (sc) The expression obtained for ⌺ can be in principle ϫ ͑ ͒ ͗cϪk ,↓ck ,↑͔͘, 35 directly inserted in Eq. ͑17͒ in order to use the resulting 2 4 ˆ (SC) ͑ ͒ or, summing over n to close the sum of 4 momenta, and Gk for correction of the gap equation 11 . However, at ␰ ϳ␧ using the spectral representation ͑9͒ for Tϭ0: the quasiparticle energies k D important there, renormal- ⌬ ization effects are negligible, and thus remains well ap- V2 ϱ 2 proximated by the result of Appendix C. ␦ ϭ ͚ ␥ ␥ ͫ ͵ d␧Im͗͗cϪ ↓c ↑ 2 k1 k2 k1 , k2 , N k1 ,k2 ,q 0 ϱ ϫ͉ ͘͘Ϫ ͵ ␧ ͗͗ ͉ ͘͘ cϪk ϩq,↓ck Ϫq,↑ d Im cϪk ,↓ ¯nk ,↑ 4. NONUNIFORM EFFECTS 2 1 0 1 2

The SPGFs considered in the previous Section describe ϱ ϫ ͵ d␧ Im͗͗cϪ ϩ ↓͉c Ϫ ↑ͬ͘͘. ͑36͒ k q, k q, the uniform self-averaging characteristics of the SC state. 0 2 1 The next important question is the behavior of fluctuations of the order parameter ͑both its amplitude and phase͒ in an Here besides the previously used SPGFs, the TPGF ͗͗cϪ ↓c ↑͉cϪ ϩ ↓c Ϫ ↑͘͘ appears. An explicit calcu- inhomogeneous system, which should be closely related to k1 , k2 , k2 q, k1 q, ϭ the breakdown of superconductivity in the overdoped re- lation of this function for the unperturbed (VL 0) SC sys- gime. A strong local suppression of d-wave order close to a tem ͑see Appendix E͒ gives the following result single ‘‘foreign’’ impurity has been predicted theoretically14 ͉ ͗͗cϪk ,↓ck ,↑ cϪk ϩq,↓ck Ϫq,↑͘͘ and observed experimentally.11 In the general case of a finite 1 2 2 1 concentration of scatterers, the local d-wave order can be 4␧⌬ ⌬ ϩ k1 k2 ¯ characterized by the operator ϭ␦ . ͑37͒ 0,q͑␧2ϪE2 ϪE2 ͒2Ϫ4E2 E2 k1 k2 k1 k2 Here E2ϭ␰2ϩ⌬2, and three other terms in the numerator are V k k k i͑kϪkЈ͒ n not mutually odd in ␥ and ␥ and thus do not contribute to ⍀ ϭ ͚ ␥ e • cϪ ↓c ↑ , ͑33͒ k1 k2 n k k, kЈ, 2 N k,kЈ ␦ . Then it is easy to see that the resulting contribution to Eq. ͑36͒ from the M-diagonal (qϭ0) and N-nondiagonal ͑ ͒ such that its mean value ͑generally complex͒ defines the uni- TPGFs by Eq. 37 exactly cancels with that from form gap parameter: NϪ1⌺͗⍀ ͘ϭNϪ1⌺␥ ⌬ ϭ⌬͑Ref. 14͒. the SPGFs, which confirms the uniformity of d-wave order n k k 2 n k in this case. The most important contribution to ␦ at Þ On the other hand, it is natural to characterize local fluctua- VL 0 comes from two consecutive scattering processes in ⍀ ͑ Þ tions of the order parameter by the variance of n identified the left-hand sides of an N- and M-nondiagonal (q 0) with the variance of the gap parameter͒: TPGF, first cϪ ↓c ↑→cϪ ϩ ↓c ↑ and then k1 , k2 , k1 q, k2 , cϪ ϩ ↓c ↑→cϪ ϩ ↓c Ϫ ↑ ͑or vice versa, Fig. 2͒,on k1 q, k2 , k1 q, k2 q, the same scattering center p. It is linear in c, while the con- ϱ 1 tribution from SPGFs in this order, Ϫ2⌬V͐ d␧ Im Tr␶ˆ ⌺ˆ , ␦2ϭ ͑͗⍀2͘Ϫ͗⍀ ͘2͒ ͑ ͒ 0 1 ͚ n n . 34 N n is zero, accordingly to the N-diagonal form of ⌺ˆ established 772 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Loktev and Yu. G. Pogorelov in the previous Section. Hence we generally estimate the 2000/PO, and by the Swiss National Science Foundation ͉␦͉ϭ⌬ͱ ͑ ͒ ͑ ͒ variance of the gap to grow with c as c/c1, where SCOPES Project 7UKPJ062150.00/1 . One of us V.M.L. ϳ 2 c1 (V/VL) defines the upper critical concentration for SC expresses his gratitude for the warm hospitality he felt at at Tϭ0. visiting Centro de Fı´sica do Porto and Institute of Physics of A more detailed analysis, resolving amplitude and phase University of Neuchatel. fluctuations, can be done in a similar way, but considering separately the two operators: APPENDIX A ⍀ ϩ⍀ϩ ⍀ Ϫ⍀ϩ n n n n ⍀ ϩϭ , and ⍀ Ϫϭ , For the normal phase, we consider the dispersion rela- n, 2 n, 2i ␧ ϭ Ϫ ϩ tion k 4t 2t(cos akx cos aky) and relate the dopant con- ⍀ centration c to the number ͑per unit cell͒ of occupied states such that their mean values ͗ n,Ϯ͘ lead to real and imagi- nary parts of the order parameter, and constructing the cor- below ␮: responding variances: 2a2 4␮ ϭ ͵ ϭ ͩ Ϫ ͪ c 2 dk arccos 1 arccos 1 ͑ ␲͒ ␧ р␮ 2 2 2 2 k W ␦ ϭ ͚ ͑͗⍀ ͘Ϫ͗⍀ Ϯ͘ ͒. Ϯ N n,Ϯ n, n 4␮ 2a2 ϫͩ Ϫ Ϫ ͪ ϭ ͵ ͵ 2 cos akx 2 dkx dky This approach should be particularly important upon ex- W ␲ 0 0 tension of the theory to finite temperatures, in order to estab- lish the dominant type of fluctuations due to the static disor- 2 1 du 1 dv ϭ ͵ ͵ der, responsible for breakdown of SC order at TϭT , and its ␲2 2 2 c 1Ϫ4␮/Wͱ1Ϫu 2Ϫ4␮/WϪuͱ1Ϫv possible role in the persistence of a pseudogap in the density ␳ Ͼ of states at T Tc . 2 4␮ ϭ Fͩ ͪ . ͑A1͒ ␲2 W 5. CONCLUSIONS Here the dependence of the integral

The analysis presented above shows that the disordered 1 arccos͑2ϪxϪu͒du structure of doped HTSC systems is crucial for many of their F͑x͒ϭ ͵ 1Ϫx ͱ1Ϫu2 characteristic properties and for the existence of SC order itself. The interplay between doping and disorder effects can is very well approximated by a simple linear function F(x) be briefly summarized as follows. Superconductivity sets in Ϸ5x/3 in the whole physically important range 0рx ϭ ␮ р ͑ Ϸ ͒ ␮ with the metallization of the system, at a critical concentra- 4 /W xmax where xmax 0.6 corresponds to max ϳ Ϸ Ϸ ͒ tion c c0 resulting from the competition between the ki- 0.15 W at the maximum physical doping cmax 0.2 . Then netic energy of the charge carriers in the regular lattice and we readily arrive at the estimate cited above: their attraction to random dopant centers. The uniform 3␲2 3cW d-wave order parameter ⌬ increases with increasing number ␮Ϸ cWϷ . of charge carriers as ⌬ϳͱc ͑Ref. 6͒ and saturates at a certain 40 4 ϳ␧ ␮Ͼ␧ optimum doping copt D /W, when the relation D comes to hold. With further increase in c, the increasing local APPENDIX B ⌬ fluctuations of bring it to collapse at some upper critical For the uniform SC system with d-wave gap, we perform ϳ 2 concentration c1 (V/VL) , resulting from the competition integration over the Brillouin zone with the parametrization between the pairing and scattering potentials. This picture is Ϫ ϭ Ϫ1 ␳ ␰ ϩ⌬Ϫ1␩ ϫ ϭ Ϯ & Ϯ & k ki a ( 0 ei ei ez), ei ( 1/ , 1/ ,0), quantitatively satisfied with a very natural choice of param- e ϭ(0,0,1), near 4 nodal points k ϭarccos(1Ϫ␮/W)e of the ϳ ϳ ϳ ␧ ϳ z i i eters W 2 eV, VL 0.5 eV, V 0.22 eV, D 0.2 eV, giv- gap function ⌬ ϭ⌬␥ . This integration for G(0) becomes ϳ ϳ k k ing plausible estimates: c0 5%, copt 15%, Tc,max ⌬ ␧ ϳ100 K, c ϳ20%. Our forthcoming work should also ͑ ͒ 1 ͑ ͒ 1 D 1 Gˆ 0 ϭ ͚ Gˆ 0 ϭ␳ ͫ ͵ d␩ ͵ d␰gˆ ͑␧,␰,␩͒ k 0 ⌬ Ϫ⌬ Ϫ␧ specify such important aspects, left beyond the scope of this N k 2 D paper, as the disorder effects on the cusp of density of states Ϫ␧ Ϫ␮ ␳͑␧͒ ␧ϭ⌬ D W at , the matching conditions between the self- ϩ ͵ d␰gˆ ͑␧,␰,0͒ϩ ͵ d␰gˆ ͑␧,␰,0͒ͬ, ͑B1͒ Ϫ␮ ␧ consistent and GE descriptions of SPGF, the exact numerical D coefficient for the critical value c , etc. And, of course, it is 1 ͑ ˆ (0) ͑ ͒͒ of fundamental interest to extend the present self-consistent where we have defined the matrix function cf. Gk in 5 treatment to the case of finite temperatures up to Tc , in order ␧ϩ␰␶ˆ ϩ␩␶ˆ ͑␧ ␰ ␩͒ϭ͑␧Ϫ␰␶ˆ Ϫ␩␶ˆ ͒Ϫ1ϭ 3 1 to obtain a quantitative estimate for the bell-like Tc(c) gˆ , , 3 1 ␧2Ϫ␰2Ϫ␩2 . Ͼ shape, and further to T Tc , to study the role of doping disorder versus d-wave SC coupling in the formation and The integration in ␰ ͑normal to the Fermi surface͒ in Eq. ͑ ϳ ͒ ͑ ͒ Ϫ␧ ␧ subsequent merging at c copt of the pseudogap in the nor- B1 treats the BCS shell, ͓ D , D͔, separately from the Ϫ␮ Ϫ␧ ␧ Ϫ␮ mal density of states. out-of-shell segments, ͓ , D͔ and ͓ D ,W ͔, where ⌬ ͑ ␥ ͒ This work was supported in parts by the Portuguese Pro- the gap parameter becomes zero together with k and no ␩ϭ⌬␥ ͑ ͒ gram PRAXIS XXI under Grant 2/2.1/FIS/302/94, by the integration over k is needed. Equation B1 permits NATO Program OUTREACH under Grant CP/UN/19/C/ one to define explicitly the coefficient functions gi in the Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Loktev and Yu. G. Pogorelov 773

ˆ (0)ϭ ϩ ␶ ϩ ␶ 2ϭ␧2 ␧ 3 ϱ general form G g0 g1 ˆ 1 g3 ˆ 3 . Let us denote z ͩ D ͪ ͵ arcsinh ydy Ϫ␩2 8 4 , and then the shell contribution to g0 results from the ⌬ ␧ /⌬ y integral: D ␧ ␧ ͱ␧2 ϩ⌬2 ␧ 3 ⌬ ␧ ␰ ␲ 8 D D D 4 D D d 2 z 2 2i ϭ ϩ Ϫ ͩ ͪ ͵ ϭ ͩ ϩ ␲ͪ Ϸ ϩ ͑ ͒ arcsinh 4 2 arcsinh 2 2 arctanh i , B2 3 ⌬ 3⌬ 3 ⌬ ␧ Ϫ␧ Ϫ␰ ␧ ␧ D Dz z D D z ␧ 8 2 D 1 which is followed for its last term by: Ϸ ͩ ln ϩ ͪ . 3 ⌬ 3 ⌬ d␩ ⌬ ͵ ϭ2 arccos . ͑B3͒ Ϫ⌬ͱ␧2Ϫ␩2 ␧ The out-of-shell contributions are: The equation for critical temperature Tc , corresponding to ⌬ϭ0, in this case reads Ϫ␧ D d␰ 1 ␧ ␧ 1 1 ͵ ϭ ͩ arctanh Ϫarctanh ͪ Ϸ Ϫ 2 2 ␧ ␮ ␧ ␮ ␧ Ϫ␮ ␧ Ϫ␰ D D ␧ ϱ ͑B4͒ 1 16 D ␰ d␧ ϭ ͵ ␰ ͩ ͪ ͵ ␭ ␲ Im d tanh ␧2Ϫ␰2Ϫ and 3 0 2kBTc 0 i0

Ϫ␮ ␥ ␧ W d␰ 1 1 8 2 E D ͵ Ϸ Ϫ . ͑B5͒ Ϸ lnͩ ͪ , ␧ 2 2 Ϫ␮ ␧ 3 ␲k T D ␧ Ϫ␰ W D B c

To find g1 and g3 , we use the obvious equations ␧ ␥ Ϸ D ␰ d␰ 1 with the Euler constant E 1.781. Hence for the d-wave ͵ ϭ ͵ ␩ d␩ϭ0 Ϫ␧ 2 2 Ϫ case the effective coupling constant is ˜␭ϭ8␭/3, which is 8/3 Dz Ϫ␰ 1 times the ‘‘Hamiltonian’’ value ␭, which can serve as one and more explanation for high Tc itself. Also, the ratio r ϭ ⌬ ͑ ͒ Ϫ␧ Ϫ␮ 2 2 2 /kBTc here turns out to be exp 1/3 times the common D ␰ d␰ W ␰ d␰ 1 ␮ Ϫ␧ ͵ ϩ ͵ ϭ ϭ ␲ ␥ Ϸ 2 2 ln 2 2 s-wave BCS value rBCS 2 / E 3.52, reaching values as Ϫ␮ ␧2Ϫ␰2 ␧ ␧ Ϫ␰ 2 ͑WϪ␮͒ Ϫ␧ Ϸ D high as rd 4.92. In turn, this means that, for other condi- ͑ ␳ ͒ ␮ tions say, 0 and V equal, the s-condensate turns out more Ϸ ͑ ͒ ln Ϫ␮ . B6 stable against thermal fluctuations and requires a higher Tc to W destroy it than the d-condensate. Of course, this is directly Summing up Eqs. ͑B2͒–͑B6͒ we obtain related to the absence of a gap in the latter case, permitting quasiparticles to exist at any TϽT . W ␲ ⌬ i␲ c g ϭ␧␳ ͫ Ϫ ͩ arccosh Ϫ ͪͬ; 0 0 ␮͑WϪ␮͒ ⌬ ␧ 2 ͑ ͒ ϭ ϭ␳ ␮ in accordance with Eq. 15 , g1 0 and g3 0 ln͓ /(W Ϫ␮)͔. APPENDIX D

ˆ (SC)ϭ APPENDIX C Calculation of the self-consistent SPGF G G0 ϩG ␶ˆ ϩG ␶ˆ generalizes the scheme of Appendix B: ͑ ͒ ϭ 1 1 3 3 The gap equation 11 for a uniform (VL 0) d-wave system at Tϭ0 transforms into:

⌬ ␧ ϱ ␧ ⌬ ␧ 1 16 D d 1 1 D ϭ ͵ ␩2 ␩ ͵ ␰ ͵ ͑SC͒ ͑SC͒ 3 Im d d 2 2 2 , Gˆ ϭ ͚ Gˆ ϭ␳ ͫ ͵ d␩ ͵ d␰gˆ ͑␧ ␭ ␲⌬ ␧ Ϫ␰ Ϫ␩ Ϫi0 k 0 0 0 0 N k 2⌬ Ϫ⌬ Ϫ␧ ͑C1͒ D Ϫ␧ with coupling constant ␭ϵV␳ . The quantity sought is the Ϫ⌺ ␰ϩ⌺ ␩ϩ⌺ ͒ϩ ͵ D ␰ ͑␧Ϫ⌺ ␰ 0 0 , 3 , 1 d gˆ 0 , gap amplitude ⌬. On the right-hand side of Eq. ͑C1͒ we Ϫ␮ ␧ perform the elementary integration over , using the relation Ϫ␮ Ϫ W Im(xϪi0) 1ϭ␲␦(x): ϩ⌺ ,⌺ ͒ϩ ͵ d␰gˆ ͑␧Ϫ⌺ ,␰ϩ⌺ ,⌺ ͒ͬ. 3 1 0 3 1 ␧ ⌬ ␧ ϱ D 16 D d␧ ͵ ␩2 ␩ ͵ ␧ ͵ ͑ ͒ Im d d 2 2 2 D1 ␲ 0 0 0 ␧ Ϫ␰ Ϫ␩ Ϫi0

⌬ ␧ D d␰ ϭ8 ͵ ␩2d␩ ͵ . ͑C2͒ 2ϭ ␧Ϫ⌺ 2Ϫ ␩ϩ⌺ 2 ϭ ϩ ͱ␰2ϩ␩2 Next we set z ( 0) ( 1) and z r ip, and 0 0 ␰ ϭ␰ϩ⌺ ⌺ ͓ pass from to x 3 , where 3 can be taken real as Then, integrating out in ␰ and passing from ␩ to y seen, e.g., from the final result ͑28͔͒. Then the analog to Eq. ϭប␧ ␩ ͑ ͒ ͑ ͒ D / , we present Eq. C2 as B2 is: 774 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Loktev and Yu. G. Pogorelov

␧ ϩ⌺ and ͵ D 3 dx 2 2 Ϫ␧ ϩ⌺ z Ϫx D 3 W G ϭ⌺ ␳ ͫ ␧ ϩ⌺ ϩ ͒2ϩ 2 ␧ Ϫ⌺ ϩ ͒2ϩ 2 1 1 0 ␮͑ Ϫ␮͒ 1 1 ͓͑ D 3 r p ͔͓͑ D 3 r p ͔ W ϭ ͭ ln 2z 2 ͓͑␧ Ϫ⌺ Ϫr͒2ϩp2͔͓͑␧ ϩ⌺ Ϫr͒2ϩp2͔ D 3 D 3 2i␲ ␧ ϩ⌺ ϩ ␧ ϩ⌺ Ϫ Ϫ ͬ. D 3 r D 3 r ͱ͑␧Ϫ⌺ ͒2Ϫ͑⌬ϩ⌺ ͒2ϩͱ͑␧Ϫ⌺ ͒2Ϫ͑⌬Ϫ⌺ ͒2 Ϫiͫarctan ϩarctan 0 1 0 1 p p At least, G is obtained after ␧ Ϫ⌺ ϩ ␧ Ϫ⌺ Ϫ 3 D 3 r D 3 r ϩ ϩ ͬͮ ␧ ϩ⌺ ͑␧ Ϫ⌺ ͒2Ϫ 2 arctan arctan D 3 xdx 1 z p p ͵ ϭ D 3 2 2 ln 2 2 , Ϫ␧ ϩ⌺ z Ϫx 2 ͑␧ ϩ⌺ ͒ Ϫz ␲ϩ ␧ D 3 D 3 2 2p/ D Ϸ Ϫi , ͑D2͒ Ϫ␮ϩ⌺ 2 2 ␧ W 3 xdx 1 ͑␧ ϩ⌺ ͒ Ϫz D z ͵ ϭ D 3 2Ϫ 2 ln Ϫ␮ϩ⌺ ͒2Ϫ 2 , ␧ ϩ⌺ z x 2 ͑W 3 z ␧ D 3 where the small term 2p/ D can be safely dropped. The next Ϫ␧ ϩ⌺ 2 2 integration, over yϭ␩ϩ⌺ , is done only on the i␲/z term D 3 xdx 1 ͑␮Ϫ⌺ ͒ Ϫz 1 ͵ ϭ 3 2 2 ln 2 2 , accordingly to Ϫ␮ϩ⌺ Ϫ ͑␧ Ϫ⌺ ͒ Ϫ 3 z x 2 D 3 z

⌬ϩ⌺ 2 ⌬ϩ⌺ 1 1 ⌬ 1 dy ͵ 2 ϭ ϩ⌺2 ͵ ⌬ y dy 1, Ϫ⌬ϩ⌺ ͱ ␧Ϫ⌺ ͒2Ϫ 2 2 Ϫ⌬ϩ⌺ 3 1 ͑ 0 y 1 ⌬ϩ⌺ ⌬Ϫ⌺ in the form 1 1 ϭarccos ϩarccos , ͑D3͒ ␧Ϫ⌺ ␧Ϫ⌺ 2 2 2 0 0 ␮ ͑␧Ϫ⌺ ͒ Ϫ⌬ /3Ϫ⌺ ϭ ͭ ϩ ⌺ 0 1ͮ ͑ ͒ G3 gr0 ln 2 3 3 . D8 WϪ␮ ␧⌬ which is relevant for G0 , supplemented with

⌬ϩ⌺ APPENDIX E ͵ 1 ydy 2 2 Ϫ⌬ϩ⌺ ␧Ϫ⌺ Ϫ ͑ ͒ 1 0 y Search for solutions of Eq. 5 in the form of GE consists in consecutive iterations of its right-hand side, separating 4⌬⌺ ϭϪ 1 systematically the GFs already present in previous ͱ ␧Ϫ⌺ ͒2Ϫ ⌬ϩ⌺ ͒2ϩͱ ␧Ϫ⌺ ͒2Ϫ ⌬Ϫ⌺ ͒2 27 ͑ 0 ͑ 1 ͑ 0 ͑ 1 iterations. Let us start from the M-diagonal SPGF Gk , and ͑ ͒ then the iteration sequence begins with singling out the scat- D4 ˆ Þ tering term with Gk itself from those with GkЈ,k , kЈ k: for G1 . The out-of-shell integration of the components men- ͑ ͒ ͑ ͒ 1 ͑ Ϫ ͒ ˆ ˆ ͑ ͒ ˆ ϭ ˆ 0 ϩ ˆ 0 i k kЈ pVGkЈ,kϭ ˆ 0 tioned is much easier, giving: Gk Gk Gk ͚ e • Gk N kЈ,p ␮Ϫ⌺ 3 dx ͵ ͑ ͒ ͑ ͒ 1 ͑ Ϫ Ј͒ 2 2 ϩ ˆ 0 ˆ ˆ ϩ ˆ 0 ˆ i k k p ˆ ␧ Ϫ⌺ ͑␧Ϫ⌺ ͒ Ϫ cG VGk G V ͚ e • GkЈ,k . D 3 0 x k k N kЈÞk,p 1 ␧Ϫ⌺ gϪ⌺ ͑ ͒ ϭ 0 Ϫ 0 E1 ␧Ϫ⌺ ͩ arctanh ␮Ϫ⌺ arctanh ␧ Ϫ⌺ ͪ 0 3 D 3 ˆ ͑ ͒ Then for each GkЈ,k we again write down Eq. 5 and sepa- 1 1 ˆ ˆ Ϸ Ϫ ͑ ͒ rate the scattering terms with Gk and GkЈ,k in their right- ␮Ϫ⌺ ␧ Ϫ⌺ , D5 3 D 3 hand sides:

͑0͒ 1 ͑ ЈϪ Љ͒ and ˆ ϭ ˆ ˆ i k k •p ˆ GkЈ,k GkЈ V ͚ e GkЉ,k N kЉ,pЈ Ϫ␮Ϫ⌺ W 3 dx 1 1 ͵ Ϸ Ϫ ͑ ͒ 1 2 . D6 ͑0͒ ͑0͒ ͑ ЈϪ ͒ Ј ␧ Ϫ⌺ ͑␧Ϫ⌺ ͒Ϫx WϪ␮Ϫ⌺ ␧ Ϫ⌺ ϭcGˆ Vˆ Gˆ ϩGˆ Vˆ ei k k •p Gˆ D 3 0 3 D 3 kЈ kЈ,k kЈ N k Here ⌺ can also be disregarded beside W, ␮, ␧ , and then 3 D ͑0͒ 1 ͑ ЈϪ ͒ Ј Ϫ ␧ ͑ ͒ ϩ ˆ ˆ ͚ i k k •p ˆ the two 1/ D terms cancel with that from Eq. D2 . Now, GkЈ V e Gk N ЈÞ combining Eqs. ͑D2͒–͑D6͒, we obtain p p

͑0͒ 1 ͑ ЈϪ Љ͒ Ј ␲ ⌬ϩ⌺ ϩGˆ Vˆ ͚ ei k k •p Gˆ . ͑E2͒ W 1 kЈ kЉ,k ϭ͑␧Ϫ⌺ ͒␳ ͫ Ϫ ͩ N kЉÞk,kЈ;pЈ G0 0 0 ␮ Ϫ␮͒ ⌬ arccosh ␧Ϫ⌺ ͑W 2 0 Note that the pЈϭp term which gives the phase factor ⌬Ϫ⌺ ЈϪ ϩ 1Ϫ ␲ ͑ ͒ ei(k k)•p in the right-hand side of Eq. ͑E2͒, coherent to that arccosh ␧Ϫ⌺ i ͪͬ, D7 0 already figured in the last sum in Eq. ͑E1͒, is explicitly sepa- Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Loktev and Yu. G. Pogorelov 775

ЈϪ rated from incoherent ones, ei(k k)•p , pЈÞp ͑but there will In the presence of scattering, we consider only the be no such separation when doing the 1st iteration of Eq. ͑5͒ M-nondiagonal (qÞ0) case, and then Eq. ͑E4͒ turns into for the M-nondiagonal SPGF Gˆ itself͒. Continuing the kЈ,k 1 Ϫ Bˆ ͑␰ ,␰ ,⌬ ,⌬ ͒f ϭϪ ͚ ͑e iq•pAˆ f Јϩeiq•pAˆ f Ј͒, sequence, we shall explicitly collect the terms with the initial 1 2 1 2 N 1 1 2 2 ͑ ͒ p function Gk , resulting from: i all multiple scatterings on ͑E5͒ the same site p, and ͑ii͒ on the same pair of sites p and pЈ Þp. Then the summing of ͑i͒ in p produces the first term of where the vectors of ‘‘single scattered’’ TPGFs are

GE, and, if the pair processes are neglected, it will coincide ͗͗cϪ ϩ ↓c ↑͉cϪ ϩ ↓c Ϫ ↑͘͘ 30 k1 q, k2 , k2 q, k1 q, with the well-known CPA result. The second term of the ͗͗ ϩ ͉ ͘͘ ͒ ЈÞ ck Ϫq,↑ck ,↑ cϪk ϩq,↓ck Ϫq,↑ GE, obtained by summing of ii in p,p p, contains inter- Јϭ 1 2 2 1 f 1 ϩ ͉ , ˆ ͩ ͗͗cϪk ϩq,↓cϪ ↓ cϪk ϩq,↓ck Ϫq,↑ͪ͘͘ action matrices ApЈ,p generated by multiply scattered func- 1 k2 , 2 1 ϩ ϩ tions Gˆ , kЈÞk etc. ͑including their own renormaliza- ͗͗c Ϫ ↑cϪ ↓͉cϪ ϩ ↓c Ϫ ↑͘͘ kЈ,k k1 q, k2 , k2 q, k1 q, ͒ ˆ tion . For instance, the iteration for a function GkЉ,k with ͗͗cϪ ↓c Ϫ ↑͉cϪ ϩ ↓c Ϫ ↑͘͘ kЉÞk,kЈ in the last term in Eq. ͑E2͒ will give: k1 , k2 q, k2 q, k1 q, ϩ ͗͗c ↑c Ϫ ↑͉cϪ ϩ ↓c Ϫ ↑͘͘ k1 , k2 q, k2 q, k1 q, ЉϪ ٞ͒ Љ Јϭ ͑ 1 0͒͑ ˆ ϭ ˆ ˆ ͚ i k k •p ˆ f 2 ϩ ͉ , GkЉ,k GkЉ V e Gkٞ,k ͩ ͗͗cϪk ,↓cϪ ϩ ↓ cϪk ϩq,↓ck Ϫq,↑ͪ͘͘ N 1 k2 q, 2 1 kЉ,pЉ ϩ ϩ ͗͗c ↑cϪ ϩ ↓͉cϪ ϩ ↓c Ϫ ↑͘͘ k1 , k2 q, k2 q, k1 q, ͑0͒ 1 ͑ ЉϪ ͒ ͑0͒ 1 ͑ ЉϪ ͒ Ј ϭGˆ Vˆ ei k k •pGˆ ϩGˆ Vˆ ei k k •p Gˆ kЉ N k kЉ N k and the 4ϫ4 matrices: ˆ ˆ ␶ ϩ ˆ ˆ V 0 V ˆ 3 0 terms with GkЈ,k and GkЉ,k ˆ ϭͩ ͪ ˆ ϭͩ ͪ A1 ,A2 . 0 Vˆ 0 ϪVˆ ␶ˆ ϩ ˆ Þ ͒ ͑ ͒ 3 terms with Gkٞ,k ͑kٞ k,kЈ,kЉ . E3 Ј Next, the equations of motion for f 1,2 : Consequently, the GE for Gˆ acquires the form given by k 1 Eq. ͑6͒. ˆ ␰Ј ␰ ⌬Ј ⌬ ͒ ЈϭϪ Ϫiq p ˆ B͑ , 2 , , 2 f e • A1 f Љ, Now turn to the TPGF 1 1 1 N ͑ ͒ ͉ E6 ͗͗cϪk ,↓ck ,↑ cϪk ϩq,↓ck Ϫq,↑͘͘, 1 Ϫ 1 2 2 1 Bˆ ͑␰ ,␰Ј ,⌬ ,⌬Ј͒f ЈϭϪ e iq•pAˆ f Љ, 1 2 1 2 2 N 2 beginning from the equations of motion in the absence of scattering, which develop into a 4ϫ4 matrix form: with ␰Јϵ␰ ␰Јϵ␰ ⌬Јϵ⌬ ⌬Јϵ⌬ ˆ ␰ ␰ ⌬ ⌬ ͒ ϭ␦ 1 k Ϫq , 2 k Ϫq , 1 k Ϫq , 2 k Ϫq , B͑ 1 , 2 , 1 , 2 f 0,qd, 1 2 1 2 contain the ‘‘doubly scattered’’ TPGF f Љ Bˆ ͑␰ ,␰ ,⌬ ,⌬ ͒ 1 2 1 2 ϵ͗͗cϪ ϩ ↓c Ϫ ↑͉cϪ ϩ ↓c Ϫ ↑͘͘, which is already k1 q, k2 q, k2 q, k1 q, ␧Ϫ␰ Ϫ␰ ⌬ ⌬ 1 2 1 2 0 M-diagonal and hence can be taken just in the form of Eq. ͑ ͒ ⌬ ␧ϩ␰ Ϫ␰ 0 ⌬ 37 . Finally, the solution ϭ 1 1 2 2 ͩ ⌬ ␧Ϫ␰ ϩ␰ ⌬ ͪ f ϭBˆ Ϫ1͑␰ ,␰ ,⌬ ,⌬ ͒ 2 0 1 2 1 1 2 1 2 ⌬ ⌬ ␧ϩ␰ ϩ␰ 0 2 1 1 2 c ϫͭ ␦ ϩ ͓Aˆ Bˆ Ϫ1͑␰Ј ,␰ ,⌬Ј ,⌬ ͒Aˆ ͑E4͒ 0,q N 1 1 2 1 2 1 with 4-vectors: ϩ ˆ ˆ Ϫ1͑␰ ␰Ј ⌬ ⌬Ј͒ ˆ ͔ ˆ Ϫ1͑␰Ј ␰Ј ⌬Ј ⌬Ј͒ͮ A2B 1 , 2 , 1 , 2 A2 B 1 , 2 , 1 , 2 d, ͗͗cϪ ↓c ↑͉cϪ ϩ ↓c Ϫ ↑͘͘ k1 , k2 , k2 q, k1 q, ͗͗ ϩ ͉ ͘͘ ͑E7͒ ck ,↑ck ,↑ cϪk ϩq,↓ck Ϫq,↑ ϭ 1 2 2 1 f ϩ ͉ , defines the contribution ϳc(V ⌬/V)2 to ␦2, Eq. ͑36͒, ͩ ͗͗cϪk ,↓cϪ ↓ cϪk ϩq,↓ck Ϫq,↑ͪ͘͘ L 1 k2 , 2 1 ⌬2 ⌬ ⌬ ϩ ϩ ͉ the factor being due to 1 , 2-odd terms ͗͗c ↑cϪ ↓ cϪk ϩq,↓ck Ϫq,↑͘͘ Ϫ k1 , k2 , 2 1 ˆ 1 ␰Ј ␰ ⌬ ⌬ 2 from B ( 1 , 2 , 1 , 2) and 1/V due to the ⌬ ⌬ 2⌬ ⌬ dominanting, zeroth order in 1 , 2 , terms 1 2 Ϫ Ϫ ⌬ from Bˆ 1(␰Ј ,␰ ,⌬Ј ,⌬Ј)Bˆ 1(␰Ј ,␰Ј ,⌬Ј ,⌬Ј) and 1 1 2 1 2 1 2 1 2 dϭͩ ͪ , Bˆ Ϫ1(␰ ,␰Ј ,⌬ ,⌬Ј)Bˆ Ϫ1(␰Ј ,␰Ј ,⌬Ј ,⌬Ј). ⌬ 2 1 2 1 2 1 2 1 2 2 *E-mail: [email protected] and ␰ ϵ␰ , ␰ ϵ␰ , ⌬ ϵ⌬ , ⌬ ϵ⌬ . The solution to 1͒ 1 k1 2 k2 1 k1 2 k2 Here we don’t consider the possible formation of stripe structures, where Eq. ͑E4͒: an ordered or disordered distribution of dopants cannot yet be confirmed by any reliable data. f ϭ␦ BϪ1͑␰ ,␰ ,⌬ ,⌬ ͒d, 2͒However, it is known that ␮ can differ substantially from the Fermi energy 0,q 1 2 1 2 ␧ ͑ ͒ F in the limit of very low doping see, e.g., Refs. 4 and 6 . gives just the result of Eq. ͑37͒ for the 1st component of f. 3͒So, we do not take into account next-neighbor hoppings. 776 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Loktev and Yu. G. Pogorelov

4͒ Ͼ 15 This approximation is actually justified by the fact that for c c0 the Fermi A. M. Nersesyan, A. Tsvelik, and F. Wenger, Nucl. Phys. B 438,561 ␧ ͑ ͒ level F of the metallic phase is well higher than the conduction band 1995 . 16 edge, and one can hardly suppose the existence of local pairs and the A. Ghosal, M. Randeria, and N. Triverdi, Preprint cond-mat/0012304 ␮Ͻ␧ ͑2000͒; Phys. Rev. B 65, 20505 ͑2001͒. related inequality F at these concentrations. Therefore, in what fol- 17 ͑ ͒ lows we do not distinguish between ␮ and ␧ . A. C. Durst and P. A. Lee, Phys. Rev. B 62, 1270 2000 . F 18 5͒Except for the special case 1/V˜ ϭ0, which corresponds to cϷ4c /3, while N. F. Mott and E. A. Davis, Electronic Processes in Non-Crystalline Ma- 0 terials, Clarendon Press, Oxford ͑1971͒. the actual consideration is for cӶc . 0 19 V. M. Loktev and Yu. G. Pogorelov, Physica C 272, 151 ͑1996͒. 20 C. C. Tsuei, J. R. Kirtley, C. C. Chi, L. S. Yu-Jahnes, A. Gupta, T. Shaw, J. Z. Sun, and M. B. Ketchen, Phys. Rev. Lett. 73, 593 ͑1994͒. 21 ͑ ͒ 1 P. W. Anderson, J. Phys. Chem. Solids 11,26͑1959͒. D. J. Scalapino, Phys. Rep. 250,329 1995 . 22 ͑ ͒ 2 A. A. Abrikosov and L. P. Gor’kov, Zh. E´ ksp. Teor. Fiz. 39, 1781 ͑1961͒ D. J. Van Harlingen, Rev. Mod. Phys. 67, 515 1995 . 23 ͑ ͒ ͓Sov. Phys. JETP 12, 1243 ͑1961͔͒. A. A. Abrikosov, Preprint cond-mat/9912394 , 1999 ; Phys. Rev. B 63, ͑ ͒ 3 Z.-X. Shen and D. S. Dessau, Phys. Rep. 253,1͑1995͒. 134518 2001 . 24 ͑ ͒ 4 V. M. Loktev, Fiz. Nizk. Temp. 22,3͑1996͓͒Low Temp. Phys. 22,3 P. W. Anderson, Phys. Rev. 102, 1492 1958 . 25 ͑ ͒ ͑1996͔͒. I. M. Lifshitz, Adv. Phys. 13, 483 1964 . 26 5 A. V. Puchkov, D. N. Basov, and T. Timusk, J. Phys.: Condens. Matter 8, I. M. Lifshitz, S. A. Gredeskul, and L. A. Pastur, Introduction to the ͑ ͒ 10049 ͑1996͒. Theory of Disordered Systems, Wiley and Sons, New York 1988 . 27 ͑ ͒ ͑ ͓͒ 6 V. M. Loktev, R. M. Quick, and S. G. Sharapov, Phys. Rep. 349,1͑2001͒. M. A. Ivanov, Fiz. Tverd. Tela Leningrad 12, 1508 1971 Sov. Phys. ͑ ͔͒ 7 D. Bonn, S. Kamal, K. Zhang, R. Liang, D. J. Baar, E. Klein, and W. N. Solid State 12, 1895 1971 . 28 Hardy, Phys. Rev. B 50, 4051 ͑1994͒. M. A. Ivanov, V. M. Loktev, and Yu. G. Pogorelov, Phys. Rep. 153,209 ͑ ͒ 8 A. Maeda, T. Yabe, S. Takebayashi, M. Hase, and K. Uchinokura, Phys. 1987 . 29 ͑ ͒ Rev. B 41,4112͑1990͒. Yu. Kagan and Ya. A. Iosilevskii, Sov. Phys. JETP 15, 182 1962 . 30 9 M. I. Salkola, A. V. Balatsky, and R. J. Schrieffer, Phys. Rev. B 55, 12 648 R. J. Elliott, J. A. Krumhansl, and P. L. Leath, Rev. Mod. Phys. 46, 465 ͑ ͒ ͑1997͒. 1974 . 31 10 Y. Fukuzumi, K. Mizuhashi, K. Takenaka, and S. Ushida, Phys. Rev. Lett. M. A. Ivanov, V. M. Loktev, and Yu. G. Pogorelov, Sov. Phys. JETP 74, ͑ ͒ 76,684͑1996͒. 317 1992 . 32 ͑ ͒ 11 S. H. Pan, E. W. Hudson, K. M. Lang, H. Esaki, S. Ushida, and J. C. L. P. Gor’kov and P. A. Kalugin, JETP Lett. 41,253 1985 . 33 Davis, Nature ͑London͒ 403, 746 ͑2000͒. M. Covington, M. Aprili, E. Paraoanu, L. H. Greene, F. Xu, J. Zhu, and ͑ ͒ 12 P. A. Lee, Phys. Rev. Lett. 73, 1887 ͑1993͒. C. A. Mirkin, Phys. Rev. Lett. 79, 277 1997 . 13 A. V. Balatsky, A. Rozengren, and B. L. Altshuler, Phys. Rev. Lett. 73, 720 ͑1994͒. This article was published in English in the original Russian journal. Repro- 14 Yu. G. Pogorelov, Solid State Commun. 95, 245 ͑1995͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Model for superconductivity in optimally doped electron cuprates T. H. Geballe* and B. Y. Moyzhes

Laboratory for Advanced Materials and Department of Applied Physics, Stanford University, Stanford, CA 94305-4045 ͑Submitted May 29, 2001͒ Fiz. Nizk. Temp. 27, 1053–1055 ͑September–October 2001͒ The very anomalous normal state properties and superconductivity of electron-doped superconducting cuprates with the T* structure are discussed. The high values of the electron and hole mobilities in the literature are explained by the electrons and holes existing in different layers. The annealing procedures necessary for superconductivity produce oxygen- vacancy negative-U pairing centers. The relative insensitivity of Tc to the annealing condition is attributed to stabilization of the chemical potential near the middle of the narrow oxygen band formed by oxygen vacancies, where the electronic entropy is large. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1401187͔

INTRODUCTION Chernik and Lykov4 of superconductivity resulting when small concentrations of Tl are substituted for Pb in PbTe. Pairing interactions which are confined only to the CuO2 Two articles review the extensive research in Russia which layers of optimally hole-doped high-temperature supercon- shows convincingly that Tl ͑and no other dopants͒ forms a ducting cuprates are believed to be insufficient to account for narrow resonant band based upon 6S0Ϫ6S2 states inside the the wide range of maximum Tc values which are found in the PbTe valence band which, when partially occupied, becomes families of superconducting cuprates which have the highest superconducting.5,6 Optimum doping of about ϳ1020 carriers ͑ ͒ Tc Ref. 1 . In the present paper we extend our discussion to per cubic centimeter results in Tc values which range up to 2 include the electron-doped superconducting cuprates. K. It is of interest that these quantities are comparable to While pairing interactions are considered to be exclu- those found in optimally oxygen-reduced SrTiO3. sively in the CuO2 layers in the most widely accepted models More experiments are needed to test the idea that nega- of cuprate superconductivity, there is certainly no theoretical tive U centers enhance the superconductivity of the layered reason why they cannot occur elsewhere. Attractive interac- cuprates. As a first step in this direction, we refer to a model tions across interfaces have been discussed many times in the 2 3 which analyzes tunneling through a barrier containing nega- past, starting with Ginzburg and Bardeen. In their models tive U impurities7 and predicts that an enhanced critical cur- artificial interfaces were considered. rent should be observed due to pair resonant tunneling. In the cuprates nature has provided layers with well- In the present paper we show that pairing due to negative matched interfaces, with the possibility of transferring charge U centers may be useful as a basis for understanding the ͑ ͒ or doping from reservoir layers which are removed from perplexing chemistry and unusual normal state properties re- the CuO2 layers. Ternary layered perovskites and structures ported in the literature of the electron-doped cuprates such as with even more-complex unit cells grown either as single ͑ ͒ ͑ ͒ NdCe 2CuO4 NCCO which have the T* structure, where a crystals or as films might be considered to be ‘‘self- separate oxygen layer exists between the rare-earth layers. organized’’ with layer sequencing that permits modulated doping. In order to account for the very high-Tc supercon- ducting transitions of the cuprates which have Hg-, Tl-, and ELECTRON DOPED CUPRATES Bi-based charge reservoir layers, we have suggested, on the basis of well-known chemistry of these elements, that those The synthesis of electron-doped superconducting cu- 6S1 cations are negative U centers. These provide an effec- prates is a two step process.8 The first step is a straightfor- tive attractive potential between electrons of opposite spin.1 ward substitution of a tetravalent cation ͑Ce or Th͒ for the The negative U center forms as a result of the cation inter- trivalent rare earth, such as Nd. The added electron which is ͑ action with surrounding atoms polarization, deformation, transferred to the CuO2 layer remains weakly localized. The magnetic, etc.͒ and internal electron correlation. In the second step involves annealing in vacuum or an inert gas, strongly ionic limit the 6S1 cations can disproportionate to which removes a small amount of oxygen ͑ϳ1or2%͒. The 6S0 and 6S2 configurations. More realistically, they would second step is essential for producing superconductivity9 and form a narrow band based upon those configurations. The cannot be understood in terms of simple doping or charge quasiparticles in this partially filled band can interact with transfer. From chemical considerations each oxygen removed the pair ordering in the CuO2 layers and become part of the leaves two electrons in the lattice. Intuitively one might ex- condensate. While at this stage there is no direct proof for pect the increased electron doping to be the same as would this idea, we believe it has merit because: ͑i͒ it is well known be obtained by the substitution of two additional Ce4ϩ for that Hg, Bi, and Tl form negative U centers in many oxides, Nd3ϩ, which results only in a small decrease in resistivity. and ͑ii͒ there is the well-documented case discovered by However, a much more remarkable change occurs. The trans-

1063-777X/2001/27(9–10)/3/$20.00777 © 2001 American Institute of Physics 778 Low Temp. Phys. 27 (99–10), September–October 2001 T. H. Geballe and B. Y. Moyzhes port behavior shows that mobile holes and electrons are cre- there are about 1020 carriers. If they were confined to single ated and that superconductivity appears below about 25 K. layers one would expect strong 2d electron-hole Coulomb The neutron diffraction data10 show no significant struc- scattering and low mobility. tural change results from the oxygen reduction ͑step 2͒. The Semimetals with high mobilities such as Sb and Bi are only difference is that a barely detectable concentration of 3-dimensional and have very large dielectric constants; both ͑ interstitial oxygen at the apical oxygen (O3) site which is of these factors increase the screening. We suggest that there not occupied in the ideal T* phase͒ is no longer detected in is a different reason for the high mobility in NCCO, namely the reduced ͑superconducting͒ single crystal. because the mobile carriers exist in different layers. The The profound consequences of removing 1% or so of electrons are confined to the CuO2 layers and the holes are oxygen from the lattice are not at all clear. Studies of confined to the oxygen layers, where the Madelung potential ͒ (Nd1.85Ce0.15 CuO4, the most intensively investigated sys- favors holes. It is significant that step 2 is only effective after tem, show that the superconducting properties are only found the long-range antiferromagnetic order is destroyed by the р in a narrow concentration range for compositions 0.15 Ce Ce doping. The electrons in the CuO2 layers are then less р0.18,9 and only after reduced oxygen annealing. After this localized and may promote increased conductivity in the treatment the transport properties suggest semimetallic be- oxygen layers. This interaction in turn should promote delo- 17 havior with low carrier concentration. For some samples the calization in the CuO2 layers. Hall effect changes sign with temperature.11 and the Seebeck 12 voltage can be of either sign. The positive Nernst effect is SUPERCONDUCTING INTERACTIONS 1 to 2 orders of magnitude larger than expected for a typical Without the oxygen vacancies there is no superconduc- Fermi liquid.12 A large positive magnetoresistance at 60 tivity. We propose that the oxygen-vacancy negative-U cen- K12,13 is observed, which on simple considerations suggests ters discussed above support superconducting pairing inter- mobilities of the order of 100 nm2/͑V s͒. Smaller changes in • actions in the oxygen layers, much as Hg, Bi, and Tl cations the resistivity itself with doping indicate that mobility do in the charge reservoir layers of the hole-doped type of changes may be more important than doping. cuprates.1 Experiments5 with Tl in PbTe wherein the hole concentration was changed by additional doping with Na, SUGGESTED MODEL which is a lower-lying acceptor level than Tl, show that, to obtain maximum Tc , the chemical potential must be close to There have been no satisfactory models proposed to ex- the middle of the narrow negative-U band, where both occu- plain the transport and superconductivity of n-type cuprates. pied and unoccupied negative U states are present. It seems apparent to us that it will be necessary to consider This position of the chemical potential is stabilized after interactions throughout the entire complex unit cell, much as annealing by thermodynamics due to the highest electron we have argued is needed to understand the hole-doped density of states being near the middle of the impurity band. 1 cuprates. For the electron-doped cuprates with the highest A somewhat related behavior has been noted in oxygen- ͑ ͒ Tc , more-definitive experiments, particularly careful mea- nonstoichiometric HgRe Ba2CuOy , where a value of Tc surements comparing single crystals and epitaxial films be- close to Tc(max) after annealing is achieved independently of fore and after step 2, are needed. At present we can offer changes in Hg and Re composition.18 The procedures for only educated guesses as to the underlying microscopic obtaining Tc in step 2, which are relatively robust and inde- physics and chemistry. pendent of the exact annealing regime, i.e., whether done in There is no doubt that in step 1 the Ce doping transfers a vacuum, or in an inert gas, or in oxygen at a more elevated electrons to the CuO2 layer. The CuO bond distance in- temperature, can reasonably be attributed to a similar ther- creases and the distance between the NdCe layer and the modynamic stabilization. oxygen layer decreases.14 CONCLUSION

NORMAL STATE PROPERTIES We have proposed a model which has the virtue of being able to explain the unusual normal state and superconducting It takes 15% Ce doping to destroy the antiferromagnetic properties of electron-doped cuprate superconductors for the long-range order, after which the transport remains in the first time. The model has unusual features, particularly that weak localization regime at low temperatures. there are mobile electrons and holes present which exist in In step 2, oxygen vacancies are created and two elec- different layers of the unit cell. Superconducting pairing in- trons remain in the lattice for each oxygen removed. It is teractions are believed to arise from negative-U center oxy- well known from the properties of reduced alkaline earth 15 gen vacancies which are formed upon removing oxygen by oxides that oxygen vacancies are never paramagnetic. We annealing. Considerably more experimental and theoretical postulate that similar behavior occurs in the reduced cu- work is required in order to prove or disprove our model. prates. That this is not unexpected follows from a model comparing the energies of singly and doubly charged impu- We are grateful for the opportunity to dedicate this short rity centers in a dielectric continuum.16 note and unconventional model to the memory of L. Shub- A maximum positive magnetoresistance is expected nikov and to acknowledge the enormous contributions to when the electron and hole conductivities are equal. This physics that he made during his shortened life. maximum is observed approximately at the same concentra- The work at Stanford has been supported in part by the ͑ ͒ tion as Tc(max) Fig. 7 of Ref. 12 . The magnitude suggests Air Force Office of Scientific Research. Low Temp. Phys. 27 (99–10), September–October 2001 T. H. Geballe and B. Y. Moyzhes 779

*E-mail: [email protected] 11 W. Jiang, S. N. Mao, X. X. Xi, X. Jiang, J. L. Peng, T. Venkatesan, C. J. Lobb, and R. L. Greene, Phys. Rev. Lett. 73, 1291 ͑1994͒. 12 1 T. H. Geballe and B. Y. Moyzhes, Physica C 341–348, 1821 ͑2000͒. P. Fournier, X. Jiang, W. Jiang, S. N. Mao, T. Venkatesan, C. J. Lobb, and 2 V. L. Ginzburg, Zh. E´ ksp. Theor. Fiz. 47, 1549 ͑1964͓͒Sov. Phys. JETP R. L. Greene, Phys. Rev. B 56, 14149 ͑1997͒. 13 20,2318͑1964͔͒. Nonsuperconducting samples show a negative magnetic resistance as ex- 3 D. Allender, J. Bray, and J. Bardeen, Phys. Rev. B 7, 1026 ͑1973͒; Phys. pected for weak localization. Rev. B 8, 4443 ͑1973͒. 14 References to original work are cited in Ref. 9. 4 I. A. Chernik and S. N. Lykov, Pis’ma Zh. E´ ksp. Teor. Fiz. 7,94͑1981͒ 15 J. W. Carson, D. F. Holcomb, and H. Ru’chardt, J. Phys. Chem. Solids 12, ͓Sov. Tech. Phys. Lett. 7,40͑1981͔͒; I. A. Chernik and S. N. Lykov, Fiz. 66 ͑1959͒. Tverd. Tela ͑Leningrad͒ 23, 1400 ͑1981͓͒Sov. Phys. Solid State 23, 817 16 B. Ya. Moizhes and S. G. Suprun, Fiz. Tverd. Tela ͑Leningrad͒ 24, 209 ͑1981͔͒. ͑1982͓͒Sov. Phys. Solid State 24,550͑1982͔͒. 5 ͑ ͒ V. I. Kaidanov and Yu. I. Ravich, Sov. Phys. Usp. 28,31 1985 . 17 We are unaware of direct experimental evidence, but do note that a me- 6 ͑ ͒ S. A. Nemov and Yu. I. Ravich, Phys. Usp. 41, 735 1998 . tallic overlayer produces a detectable increase in the T ofaMoGethin 7 V. Oganessian, S. Kivelson, T. H. Geballe, and B. Y. Moyzhes ͑to be c film. D. Ephron Thesis, Correlated Electron Tunnelling and Quantum Mo- published͒. tion of Vortices in Disordered Model Systems, Stanford University ͑1996͒; 8 Y. Tokura, H. Takagi, and S. Uchida, Nature ͑London͒ 337, 345 ͑1989͒. 9 N. Mason ͑private communication͒. Original references are given in the extensive review by J. Fontcuberta 18 and L. Fahrega, in Studies of High Temperature Superconductors, Vol. 16, A. Serquis, L. Fabrega, L. Morales, A. Basset, A. Caneiro, J. Fontcuberta, A. Narlicar ͑Ed.͒, Nova Science Publishers, Inc. ͑1996͒, p. 85. It is note- S. Short, J. P. Hodges, and J. Jorgensen, Physica C 341–348,511 worthy that similar results are obtained for a variety of annealing proce- ͑2000͒. dures. 10 A. J. Schultz, J. D. Jorgenson, J. L. Peng, and R. L. Greene, Phys. Rev. B This article was published in English in the original Russian journal. Repro- 53,5157͑1996͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

The influence of hydrogenation on superconducting properties of MgB2 A. J. Zaleski,* W. Iwasieczko, D. Kaczorowski, and H. Drulis

Institute of Low Temperature and Structure Research, PAS, Wroclaw, Poland M. Tkacz and O. J. Zogal

Institute of Physical Chemistry, PAS, Warsaw, Poland J. Klamut

International Laboratory of High Magnetic Fields and Low Temperatures, Wroclaw, Poland ͑Submitted June 5, 2001͒ Fiz. Nizk. Temp. 27, 1056–1059 ͑September–October 2001͒ In the paper we present the results of the ac susceptibility measurements of the newly discovered superconducting diboride—MgB2.©2001 American Institute of Physics. ͓DOI: 10.1063/1.1401188͔

1. INTRODUCTION Substituting magnesium with monovalent, divalent, or trivalent ions was studied by Medvedeva et al.9 Its influence 1 At the very beginning of the new millennium Akimitsu on the superconducting properties of MgB2 is connected with proved that it is still possible to discover superconductivity the electronic structure of the latter and is dominated by the with a rather high critical temperature in a compound which chemical bonding of the hexagonal, graphitelike B sheets. comes from the already well-investigated family of diborides The ␴ bonds between the boron atoms are unfilled and thus ␲ and was commercially available. After this discovery, MgB2, metallic, with more electrons in the bonds. As a result, ϭ with its critical temperature Tc 38 K, was investigated with MgB2 may be well characterized by the ionic form 2ϩ͑ ͒2Ϫ ␴ unprecedented speed. After a few months almost every char- Mg B2 . Filling the holes in the band by trivalent acteristic of the material had been determined. substitution is unfavorable for superconductivity, while some The first information about the superconducting proper- monovalent substitutions may be favorable for it. These find- ties of the compounds with the AlB2-type structure was pub- ings have been partially supported as the substitution of lished already in 1970 by Cooper et al.2 They found that trivalent aluminum degraded the superconductivity in the 10 hexagonal NbB2 or MoB2 as cast are not superconducting, material, but monovalent copper substitution only broad- but increasing the amount of boron to about 2.5 or substitut- ened the transition without changing the transition-onset 11 ing with Sc, Y, Zr, Mo, Ru, Hf, or Th for Nb or additionally value. with Al, Ti, V, Ta, or Au for Mo resulted in a superconduct- In our paper we describe the results of a different method of doping of MgB —hydrogenation. We also men- ing material with a critical temperature peaking at Tc 2 ϭ ϭ tion the results of doping TaB with hydrogen. 9.3 K for Nb1.9Y0.1B2 and at Tc 11.2 K for 2 Mo1.69Zr0.31B2. The results of another systematic search for supercon- 2. MATERIALS AND METHODS ductivity in diborides were published by Leyarovska et al.3 ͑ ϭ Investigating the magnetic susceptibility of MeB2 Me Ti, In our investigations we made use of commercially ͒ ͑ Zr, Hf, V, Nb, Ta, Cr, Mo down to 0.42 K, the authors found available, powdered MgB2 Johnson Matthey GmbH Alfa— that only NbB2 was superconducting, with a critical tempera- 98% purity͒. Hydrogen absorption was performed at a con- ϭ ture of Tc 0.62 K. MgB2 was not among the materials stant temperature under hydrogen gas pressure of 0.5–20 bar. studied. Our first attempts at a reacting temperature of about 250 °C 1 After Akimitsu’s discovery it was quite natural that resulted in decomposition of MgB2. So finally the reaction many research groups rechecked old or newly prepared di- temperature was set at 100 °C and the pressure at 20 bars. borides. And the results obtained by different groups were The concentration of hydrogen was determined volumetri- quite surprising. Kaczorowski et al.4 found a superconduct- cally by monitoring the pressure change in a calibrated, ϭ ing signal in old TaB2, with Tc 9.5 K, but did not find it for sealed volume. Under the conditions described above, the 5 TiB2, HfB2,VB2, NbB2, or ZrB2. In turn Gasparov et al. hydrogen content was not higher than about (3Ϯ1)%/f.u. In ϭ found superconductivity in ZrB2 with Tc 5.5 K, but they an attempt to increase the hydrogen uptake we increased the did not confirm its existence for TaB2 and NbB2. Following hydrogen pressure to 7 kbar, leaving the temperature equal to 6 the report by Felner that BeB2 is not superconducting, there 100 °C. In this case the uptake of hydrogen was very similar, was a paper by Young et al.7 reporting superconductivity in as measured by the mass spectrometry method, and equal to Ϸ BeB2.75, with Tc 0.7 K. Superconductivity was also about 3 %/f.u. X-ray analysis was carried out using an STOE 8 ͑ discovered in ReB2 Tc in the range 4.5–6.3 K, depending diffraction system. The material was characterized by ac ͒ ϭ on boron content and Re3 B(Tc 4.7 K). magnetic susceptibility measurements.

1063-777X/2001/27(9–10)/3/$20.00780 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 Zaleski et al. 781

FIG. 1. X-ray diffractograms of as-obtained MgB2 and hydrogenated under pressures of 20 bar and 7 kbar.

3. RESULTS AND DISCUSSION X-ray diffractograms for as-obtained and hydrogenated samples are presented in Fig. 1. It is clearly seen that both the positions of the diffraction peaks and their widths are practically the same for all the measured samples. Small changes of the peak heights are to be ascribed, according to us, to the difference between the mass and thickness of the samples rather than to the different amounts of the majority phase. In principle such a result might be expected as the FIG. 3. Comparison of the ac susceptibility measurements results for ͑᭹͒ ͑᭺͒ hydrogen uptake was rather small. Summing up—the struc- MgB2H0.03 obtained under hydrogen pressures of 20 bar and 7 kbar . ture and the volume of the phase with the MgB2-type struc- ture remained unchanged after hydrogenation. 12 According to the paper by Wan et al. the values of the level relative to the density of states ͑DOS͒. Thus the con- lattice constants are connected with the location of the Fermi stancy of the lattice constants probably means that the DOS is not changing meaningfully, either. The results of ac susceptibility measurements of MgB2 hydrogenated at 20 bar are presented in Fig. 2. For compari- son, the results for the parent MgB2 are presented in the insets. The masses of both powdered samples were very similar. So the main features of the plots are: — the critical temperatures of both pure and hydrogenated compounds differs negligibly; — the superconducting response ͑and thus the amount of superconducting phase͒ is two orders of magnitude lower for the hydrogenated sample; — the superconducting transition width is narrower for the hydratized material. Figure 3 shows a comparison between the ac suscepti- bilities of the MgB2 after hydrogenation under different hy- drogen pressures. As was mentioned above, the hydrogen uptake was very similar in the two cases despite the fact that the pressures during hydrogenation differ by 350 times. And, similarly to Fig. 2, one can see that the critical temperature remained unchanged and that there is an additional decrease of the amount of superconducting phase. The two facts—unchanged critical temperature and de- creasing ͑with hydrogenation͒ amount of superconducting phase–are rather surprising. The first fact implies that the density of states, electron–phonon interaction, and phonon spectrum remained unchanged after hydrogen uptake, ͑ ͒ ͑ ͒ whereas the decrease of the amount of superconducting FIG. 2. Real a and imaginary b parts of the ac susceptibility of MgB2 hydrogenated under a pressure of 20 bar. The insets show the ac suscepti- phase may imply that the density of states was meaningfully bility for as-obtained MgB2. changed. 782 Low Temp. Phys. 27 (99–10), September–October 2001 Zaleski et al.

So to account for our observations, in the picture of con- 4. SUMMARY ventional BCS phonon-mediated superconductivity, the in- fluence of hydrogenation on superconductivity should act in In conclusion, we have observed a meaningful decrease two opposing directions. We can analyze it using the modi- of the diamagnetic signal of hydrogenated commercial MgB2 fied McMillan equation:13 powder, while its critical temperature remained practically unchanged. The hydrogen uptake was very low ͑about 4%/ ប␻ ϩ␭͒ 0 1.04͑1 ͒ k T ϭ expͩ Ϫ ͪ , f.u. and independent on the pressure used. Such strange be- B c 1.2 ␭Ϫ␮*͑1ϩ0.62␭͒ havior can be explained by some opposite influence of the ␻ hydrogen on the DOS, phonon frequency, electron–phonon where kB is Boltzmann’s constant, 0 is the averaged pho- non frequency, ␭ is the electron–phonon coupling constant, coupling, and Coulomb screening. ␮ and * is the Coulomb pseudopotential. The authors would like to express their gratitude to Dr. Hydrogenation can influence the phonon frequency. It Marek Wolcyrz for x-ray measurements and to Dr. Tadeusz 14 was shown by Hinks et al. that the vibrations of B atoms Kopec for valuable discussions. are more strongly coupled to the electronic structure than those of Mg atoms. This may be inferred from the fact that the isotope effect for boron is much stronger than for mag- 15 nesium. Both a theoretical calculation and experimental *E-mail: [email protected] work support the view that the E2g in-plane boron stretching mode is the main source of strong electron–phonon cou- pling. This mode was found to be strongly anharmonic.16 So if hydrogen is adsorbed within the boron planes ͑where it is ͒ easy to find the proper placement for it it may strongly 1 J. Akimitsu, Symposium on Transition Metal Oxides, Sendai, January 10, influence the stretching boron mode. But it is not easy to 2001. determine what kind of influence it might have. It might be 2 A. S. Cooper, E. Corenzwit, L. D. Longinotti, B. T. Matthias, and W. H. increasing or lowering the anharmonicity or lowering or in- Zachariasen, Proc. Natl. Acad. Sci. U.S.A. 67, 313 ͑1970͒. 3 L. Leyarovska and E. Leyarovski, J. Less-Common Met. 67,249͑1979͒. creasing the phonon frequency. This way hydrogen can also 4 D. Kaczorowski, A. J. Zaleski, O. J. Zogal, and J. Klamut, cond-mat/ change ␭ or ␮*. 0103571. Absorption of hydrogen within the Mg planes also can 5 V. A. Gasparov, N. S. Sidorov, I. Izver’kova, and M. P. Kulakov, cond- mat/0104323. not be excluded. Such a placement should not change the 6 14 I. Felner, cond-mat/0102508. important modes of phonon frequencies, but it could influ- 7 D. P. Young, P. W. Adams, J. Y. Chan, and F. R. Fronczek, cond-mat/ 15 ence the ionicity of the magnesium and boron planes. This 0104063. in turn should change the position of Fermi level with re- 8 G. K. Strukova, V. F. Degtyareva, D. V. Shovkun, V. N. Zverev, V. M. spect to the ␴ boron bands and DOS. And with it also Cou- Kiiko, A. M. Ionov, and A. N. Chaika, cond-mat/0105293. 9 N. I. Medvedeva, A. I. Ivanovskii, J. E. Medvedeva, and A. J. Freeman, lomb screening and electron–phonon coupling could be cond-mat/0103157. changed. We believe that a change of ionicity of the layers 10 J. S. Slusky, N. Rogado, K. A. Regan, M. A. Hayward, P. Khalifah, T. He, should influence the value of the lattice constant c, perpen- K. Inumaru, S. M. Loureiro, M. K. Haas, H. W. Zandbergen, and R. J. ͑ ͒ ͑ ͒ dicular to the planes. As no such influence was observed in Cava, Nature London 410,343 2001 . 11 S. M. Kazakov, M. Angst, and J. Karpinski, cond-mat/0103350. our study, we are inclined toward the first possibility. 12 X. Wan, J. Dong, H. Weng, and D. Y. Xing, cond-mat/0104216. We also applied hydrogenation under low pressure and 13 W. L. McMillan, Phys. Rev. 167, 331 ͑1968͒. 14 D. G. Hinks, H. Claus, and J. D. Jorgensen, Nature ͑London͒ 411,457 temperature to the controversial compound TaB2 studied by 4 ͑2001͒. Kaczorowski et al. In this case the hydrogen uptake was 15 J. M. An and W. E. Pickett, cond-mat/0102391. equal to 36 %/f.u., much higher than for MgB2. And inspite 16 A. F. Goncharov, V. V. Struzhkin, E. Gregoryanz, J. Hu, R. J. Hemley, of such high hydrogen content, the critical temperature re- H.-K. Mao, G. Lapertot, S. L. Bud’ko, and P. C. Canfield, cond-mat/ mained unchanged and equal to about 9.5 K, but the diamag- 0104042. netic signal was an order of magnitude lower than for the This article was published in English in the original Russian journal. Repro- undoped material. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Quantum Hall effect in nonuniform two-dimensional electron systems V. B. Shikin*

Institute of Solid State Physics, Russian Academy of Sciences, 142432 Chernogolovka, Moscow District, Russia ͑Submitted February 26, 2001͒ Fiz. Nizk. Temp. 27, 1060–1068 ͑September–October 2001͒ The quantum Hall effect ͑QHE͒ is inherent to uniform two-dimensional ͑2D͒ electron systems. At the same time, the basic low-dimensional conducting systems in the presence of metallic contacts are ‘‘regularly nonuniform.’’ Nevertheless, the QHE is well observed in them. The apparent paradox is resolved in a nontrivial manner. Regularly nonuniform 2D systems under conditions of the QHE admit the existence of integer ͑incompressible͒ channels that determine the conducting properties of the 2D system. To describe such a modification of the QHE requires a special formalism that explicitly takes into account the properties of the incompressible strips. The goal of this paper is to discuss the causes for the onset of regular nonuniformity of the 2D electron density in typical low-dimensional systems and to determine the basic characteristics of the QHE for such systems in terms pertaining to the individual incompressible strips. © 2001 American Institute of Physics. ͓DOI: ͓DOI: 10.1063/1.1401189͔

It is well known that the contact of two three- where Wi is the work function of the respective metal, and e ͑ ͒ dimensional 3D metals having different work functions Wi is the elementary charge. leads to a partial transfer of electrons from one of the metals In the contact of metals A and B with open faces ͑see to the other so as to equalize their electrochemical Fig. 1͒ an electric field with potential ␸ arises in the vacuum potentials.1 The corresponding disruption of the local neu- gap between the faces AO and OB ͑Ref. 1͒: trality in good metals occurs on scales of the order of the ␪ interatomic distances, while in samples with Boltzmann sta- ␸͑␪͒ϭ␾ ͑ ͒ ab␣ , 2 tistics it occurs over a length of the order of the Debye ra- dius. It is of interest to examine the analogous problem for where ␣ is the angle between faces OA and OB. contacts between 2D and 3D conducting systems. This The electric field is given by makes it clear that in 2D systems there is no characteristic ␸ ␾ץ 1 length scale over which the local neutrality is disrupted, i.e., ϭϪ ϭ ab ϭ ␲ ␦ ͒ ͑ ͒ Er , Er 4 e n͑r 3 ␪ ␣rץ the contact-related perturbation of the electron density spans r practically the entire 2D region accessible to the charge and it falls off in inverse proportion to the distance r from transport. Below we shall refer to such a deformation of the the point O. This field is related to the surface charge density ͑ electron density as regularly nonuniform in distinction to ␦n(r) distributed along the free faces OA and OB. the possible random fluctuations of the density, the average Now let the role of ray OB be played by a 2D electron value of which over the volume of the sample is usually system and for simplicity take angles ␣ and ␤ in Fig. 1 equal ͒ assumed to be zero . In view of the sensitivity of many 2D to each other. Now the additional surface charge density problems, the QHE in particular, to the local electron density, it seems logical that the contact-related regular nonunifor- mity of the electron density will play an appreciable role in the effective formation of the Hall plateau. In this paper we give a consistent treatment of contact phenomena in 2D sys- tems and discuss the details of the QHE in the presence of contact effects.

CONTACT ELECTROSTATICS

1. In setting forth the concrete results, let us start by recalling the electrostatics for a contact of two metals having different bulk characteristics, a circumstance which is effec- tively taken into account by the introduction of a contact ␾ ͑ ͒ potential difference ab Ref. 1 , ␾ ϭ Ϫ ͑ ͒ e ab Wa Wb , 1 FIG. 1. Contact of two metallic samples having free boundaries.

1063-777X/2001/27(9–10)/7/$20.00783 © 2001 American Institute of Physics 784 Low Temp. Phys. 27 (99–10), September–October 2001 V. B. Shikin from ͑3͒ is simultaneously an addition to the uniform elec- The Fourier transformation of ͑5͒ with respect to the tron density ns of this system. In other words, the contact of variable y gives the 2D electron gas with the conducting electrodes ͑2D or 2e ϩw 3D͒ can at large distances noticeably disrupt the spatial uni- ␸͑ ͒ϭ ϭ ͵ ␦ ͑ ͒ ͑ ͉ Ϫ ͉͒ ͑ ͒ x,q const ␬ n s,q K0 q x s ds, 6 formity of the 2D electron density if such was present in the Ϫw 2D sample without the contacts. 2␲/Lр͉q͉рϩϱ, 2. The divergence of the field E ͑3͒ at small distances and the integral divergence for the total effective surface where K0(x) is the Bessel function of complex argument. charge, which also follows from ͑3͒, are removed in the natu- In the limit of small x the function K0(x) follows a ral way. The first of these singularities vanishes when various logarithmic law. Consequently, if Lӷw, then far from the corrections to the condition of thermodynamic equilibrium contacts the integral equation ͑6͒ takes the form are included in the problem. For example, if one is talking 2e2 ϩw L about an accumulation of electrons in a 2D electron gas, then ␸͑ ͒ϭ ͵ ␦ ͑ ͒ ϭ ͑ ͒ e x ␬ n s ln Ϫ ds const. 7 the divergence is removed at Bohr distances by allowance Ϫw x s ͑ for the Fermi energy in the overall balance of forces see The solution of this equation with respect to ␦n(x) has the ͒ Refs. 2–4 . If we arbitrary talking about the other sign of W, structure corresponding to depletion of the 2D system, then the result- ing equilibrium problem can contain so-called depletion lay- const ␦n͑x͒ϰ ͑8͒ ers, completely devoid of free electrons, at the ends of the ͱw2Ϫx2 2D system. This type of equilibrium has been discussed in great detail in the literature, particularly in connection with with the value of the const determined by the requirement the creation of controllable 2D systems, the geometry of 2e2 ϩw which is controlled by external fields ͑see, e.g., Refs. 2, 3, Wϭ ͵ ␦n͑s͒ln͑L/s͒ds. ␬ Ϫ 5–9͒. w Removal of the integral divergence of the charge is As a result, we have achieved by restricting the dimensions of the 2D system. The ␬W L different versions of this problem are extremely diverse. Be- ͒ϭ ϩ␦ Ӎ ϩ ϭ ␲ ͑ ͒ n͑x ns n ns , c 2 ln . 9 low we will be mainly discussing two results: a density per- ce2ͱw2Ϫx2 w turbation ␦n (x) for a quasi-1D Corbino disk, and the analo- 0 → ͑ ͒ Ӷ gous problem for an extended Hall sample in its central part, The limit q 0 used in 7 applies if w L and the in- far from the metallic contacts. vestigated x cross section of the Hall sample is far from the Ϫ ϩ Ӷ metallic contacts. Both assumptions are assumed to hold be- A Corbino disk is quasi-1D if (R1 R2)/(R1 R2) 1, low. where R1 and R2 are the outer and inner radii of the disk. In ␦ this case the uniform part n0(x) of the electron density of contact origin has been found as3 ␬ ␾ EQUILIBRIUM CHARACTERISTICS OF THE SAMPLE IN THE w ab ␦n ͑x͒ϭ , Ϫwрxрϩw. ͑4͒ QUANTUM HALL EFFECT REGIME 0 ␲2e͑w2Ϫx2͒ The formation of integer Hall strips in nonuniform 2D Here 2wϭR ϪR is the width of the 2D region between the 2 1 systems is accompanied by a redistribution of charge in the metallic ‘‘banks,’’ the x axis is directed in the radial direc- neighborhood of the integer points on the electron density tion, the origin of coordinates is at the center of the 2D profile n(x). Consequently, one of the main properties of region, ␬ is the dielectric constant of the medium, and ␾ is ab normal 2D systems—their equipotential character—is dis- the contact potential difference. In the limit a Ӷw (a is the B B turbed for states in the QHE regime. The corresponding pre- effective Bohr radius͒ approximation ͑4͒ works well far from dictions of the theory are checked by obvious experiments on the points xϭϮw. the measurement of the local electrostatic fields ͑techniques For a Hall sample the contact electrostatic problem is not utilizing the linear electrooptic effect are set forth in Refs. yet ‘‘standard,’’ and it therefore makes sense to give the de- 10–12 and the microprobe technique is described in Refs. 13 tails of its solution. The initial equation is and 14͒. The general conclusions that follow from Refs. 2e2 ϩL ϩw ␦n͑s,␴͒ 10–14 argue in favor of the existence of 2D contact phenom- ␸͑ ͒ϭ ͵ ␴ ͵ e x,y d ds 2 2 1/2 ena and, as a consequence, integer channels in the interior of ␬ ϪL Ϫw ͓͑xϪs͒ ϩ͑yϪ␴͒ ͔ 2D systems. ϭconstϭW, ͑5͒ The quantitative basis for the conclusions of this Section is provided by the results of Shklovskii and coworkers8,9 on Ϫ р рϩ Ϫ р рϩ ϵ Ϫ w x w; L y L; W Wa Wb ; the properties of equilibrium integer channels for 2D ␸͑x,y,0͒ samples with density n(x). For example, for the centralץ ␸͑x,y,zϭ0͒ϭ0,͉y͉ϾL; ϭ0,͉x͉Ͼw. channel and with a known curvature nЉ(0) of the classical zץ distribution of the electron density at its extremal point ͑the The y axis is directed along the long side of a rectangle, the point with zero first derivative nЈ(0)), the distribution ␸(x) electric potentials of the metallic contacts are both chosen over the cross section of the channel with an integer filling ϭ ␾ ␯ ϭ equal to zero, and W e ab . factor l 1,2,3,... has the form Low Temp. Phys. 27 (99–10), September–October 2001 V. B. Shikin 785

␲enЉa3 ␸͑x͒ϭϪ ͑1Ϫx2/a2͒3/2, ͑10͒ 3␬ and the width 2a is determined from ͒ 2 ϭ ␯ ͒Ϫ␯ nЉ͑0 a /4 ͓ ͑0 l͔nH , nЉ͑0͒ϭd2n͑0͒/dx2, ͑11͒ ប Ϫ c ␯͑0͒ϭn͑0͒/n , n 1ϭ␲l2 , l2 ϭ . ͑12͒ H H H H eH Here H is the magnetic field normal to the plane of the Corbino disk. The width of a strip is rather sensitive to the value of the magnetic field, and it varies from zero upon its nucleation to the critical dimensions FIG. 2. Hall sample from Ref. 11; the arrows are a schematic indication of the current streamlines. ␬ប␻ 3 c a3 ϭ , ͑13͒ max ␲e2͉nЉ͑0͉͒ where, as the magnetic field is increased, the maximum value channels, here the relative value of ␸(x) falls off toward the of the potential ͑10͒ at the center of the channel reaches a center of the cross section. We note that the available data on ប␻ ͑ value c , after which the strip splits into two for details of the ‘‘intermediate’’ state do not give the absolute values of this process see Ref. 9͒. ␸(x) but only indicate that it is nonmonotonic. Substituting the expression for nЉ(0) that follows from Turning to a discussion of the experiments, let us begin ͑4͒ into the definition ͑13͒,weget with the results presented in Ref. 11, which attest to the ␸ 3␲ប␻ nonuniformity of the distribution (x) over the cross section ͑ ͒3ϭ c ͑ ͒ of the Hall sample under the conditions of the QHE. Unfor- amax /w ␾ . 14 2e ab tunately, the data of Ref. 11 contain an uncertainty that in- For a Hall sample with n(x) given by Eq. ͑9͒ we have terferes with their interpretation: in Fig. 2, taken from Ref. 11, one can clearly see an asymmetry which is apparently 3cប␻ ͑ ͒3ϭ c ͑ ͒ due to the influence of the neighboring channels; in addition, amax /w ␲ . 15 W the position of the cross section, chosen by the authors from ͑ ͒ ͑ ͒ ប␻ у hydrodynamic considerations ͑decreasing the transport cur- Formulas 10 – 15 are valid if c W. In the oppo- ប␻ Ӷ rent through this ‘‘dead’’ channel as much as possible͒,is site, altogether typical limiting case c W it becomes pos- sible for a large number of incompressible strips to form, and located at a short and poorly determined distance from the the description of the form in ͑10͒, ͑11͒ loses meaning. An ends of this segment of complex configuration, complicating adequate formalism is needed which takes into account the the consideration of the details of ␸(x). Nevertheless, the 2 2 3/2 large number of strips and the interactions between them. function ␸(x)ϰ(1Ϫx /a ) or, more precisely, its trans- This can be done if it is taken into account that each strip has form ប␻ a potential difference c on its ‘‘banks.’’ Consequently, the 1 xϩR local value of the electrostatic potential in the problem with ␾͑x͒ϭ ͵ ␸͑s͒ds, ͑17͒ 2R Ϫ a large number of strips can be constructed in the form x R which is shown by the solid curve in Fig. 3, gives a good e␸͑x͒Ӎប␻ ␯͑x͒, ␯͑x͒ϭ␲l2 n͑x͒, ͑16͒ c H approximation of the data of Ref. 11 for the central part of ␯ ϭ ␮ where (x) is the local filling factor and lH is the magnetic the sample. Here R 50 m is the radius of the laser beam length. used in the experiments.11,12 ͑ ͒ ប␻ Ӷ According to Eq. 16 , in the limit c W the electro- The information for the Corbino disk is more ‘‘transpar- static potential reproduces on average the local behavior of ent.’’ A comparison of the data of Ref. 12 with two versions the electron density. of the distribution ␸(x)—multichannel ͑16͒ and single- Formula ͑16͒, like expressions ͑4͒ and ͑9͒, loses mean- channel ͑10͒—is presented in Figs. 4 and 5. The calculated ing near the ends Ϯw, where the excess electron density is curves ␾(x) were constructed using algorithm ͑17͒. The divergent ͑or zero͒. curves were ‘‘tied in’’ to the minimum ␾(0), and the param- Having definitions ͑10͒–͑16͒ at our disposal, let us con- eters R and the geometric dimensions contained in defini- sider the experimental data for the equilibrium distribution tions ͑10͒ and ͑16͒ were varied. For example, the solid curve ␸(x) in rectangular samples and Corbino disks.11,12 The most in Fig. 4 corresponds to the values Rϭ50 ␮m and ␻ important result of these measurements is the qualitative dif- ϭ250 ␮m. The additional lines in Fig. 4 illustrate the influ- ference in the behavior of the equilibrium value of ␸(x) over ence of the laser beam radius R on the behavior of ␾(x). In the cross section of the sample in the normal and ‘‘interme- Fig. 5 the function ␸(x)in͑10͒ was ‘‘fit’’ to the same points diate’’ states, when the system contains integer channels. In for two different values of a (Rϭ50 ␮m). the first case, as expected, ␸(x)ϭconst along the 2D system. It should be noted that the experimental points in Figs. 4 As to the ‘‘strip’’ systems containing one or several integer and 5 were obtained by scanning Figs. 1 and 2 from Ref. 12. 786 Low Temp. Phys. 27 (99–10), September–October 2001 V. B. Shikin

FIG. 5. The function ␺(x) from ͑10͒ and ͑17͒ for a Corbino disk. R FIG. 3. Distribution of the electric potential over the cross section AB ͑see ϭ50 ␮m, aϭ230 ␮m, ͑solid curve͒; Rϭ50 ␮m, aϭ250 ␮m ͑dashed Fig. 2͒ on the so-called ‘‘dead’’ end ͑there is no transport current through the curve͒. cross section͒. The solid line was calculated according to formulas ͑9͒, ͑10͒, and ͑17͒ with Rϭ50 ␮m, aϭ160 ␮m.

equilibrium conditions, when the electrochemical potential ␮ Then, using Fig. 9 from Ref. 11, which is identical to Fig. 1 is constant along the sample. Moreover, a large number of from Ref. 12, we were able to rescale the results of the scan- the experiments with nonuniform samples and, in particular, ning of Fig. 2 of Ref. 12 to the corresponding numbers for practically all the measurements of the local characteristics ␸(x). of various 2D systems ͑see Refs. 10–14͒ have been carried Summing up, we can conclude that the Hall samples out in the transport regime, when there is a ‘‘drawing’’ po- studied in Ref. 11 are well described in the single-channel tential difference V across the boundaries of the sample. approximation ͑10͒. As to the Corbino disk, the experimental H The influence of V Þ0 on the characteristics of the indi- data of Ref. 12 are better described by a multichannel distri- H vidual integer ͑incompressible͒ channels becomes a question. bution ␸(x) ͑16͒. We are talking about a change in the strip width 2a and also a shift ⌬ of their position ͑relative to the equilibrium posi- HALL TRANSPORT IN NONUNIFORM SAMPLES ͒ tion under the influence of VH . In this Section we solve the The phenomenological theory8,9 of the integer Hall chan- problem for the central incompressible channel in a nonuni- nels in nonuniform 2D electron systems was developed for form 2D electron system having a one-dimensional electron density profile n(x) which is symmetric about its center. Be- sides the one-dimensionality of the problem we shall use additional simplifications ͑as in Refs. 8 and 9͒, viz: the trans- verse dimensions 2w of the two-dimensional system are quite large ͑in comparison with the Bohr radius and the mag- netic length͒, there is no screening of the electrodes, the tem- perature is low but finite ͑in order to avoid the correlation effects that lead to the fractional quantum Hall effect͒, and the Hall potential difference VH is less than or comparable to Ӷប␻ ␻ the cyclotron energy, i.e., VH c , where c is the cyclo- tron frequency. ⌬ General results for (VH) and a(VH) are used below in an interpretation of the observed14 details of the local char- acteristics of the Hall samples with a nonuniform profile n(x) of contact origin. 1. Let us start with the electrochemical potential ␮. Un- der nonequilibrium conditions this difference is no longer constant along the 2D system. Its behavior follows from the requirements

FIG. 4. Coordinate dependence of the transform ͑4͒, ͑16͒, ͑17͒ for a Corbino disk. The solid line corresponds to parameters Rϭ50 ␮m, wϭ250 ␮m. The ٌ div jϭ0, ٌϫrot jϭ0, ͑18͒ experimental points are from Ref. 12. The degree of sensitivity of the po- • tential ͑16͒, ͑17͒ to variations of the radius R is illustrated by the two ͒ ͑ ץ ␮ץ ϭ ␮ ͑ ͒ ϭ ␮ ͑ ͒ ϭ Ϫ1␴ additional curves: R 75 m ––– , R 5 m ••••• . ji e ik / xk , 19 Low Temp. Phys. 27 (99–10), September–October 2001 V. B. Shikin 787

where ␴ is the conductivity tensor ͓in Ref. 9 ͑see Eq. ͑56͒ ␲ Ј ϩ␬ ϭ ͑ ͒ ik 2 en⌬a EH 0, 24 of that paper͒ and Ref. 15 Ohm’s law contains the electrical 2ϭϪ ␯ Ϫ Љ ͒ ϭ ͑ ͒ potential ␸ instead of ␮͔. The combination of ͑18͒ and ͑19͒ a 4͓ ⌬ l͔nH /͑n⌬ ,l 1,2,3,..., 25 leads to the equation where n(x) is the equilibrium profile of the electron density. ␮ϭ ͑ ͒ The difference between ͑25͒ and the definition of a2 from⌬ 2ٌ • 0 20 Ref. 9 lies in the fact that ␯⌬ and the derivative nЉ in our even if ␴ →0 ͑we note that the authors of Ref. 15 do not ⌬ xx case are evaluated not at zero but at the point xϭ⌬. share this point of view͒. The potential ␸(␰) is obtained from Eq. ͑23͒ by a single Using ͑20͒ and the inequality wӶL, where L is the integration. Additionally taking into account ͑24͒ and ͑25͒, length of the Hall sample, it is easy to see that in the pres- ␮ we have ence of a Hall voltage VH the function (x) for a sample far ␲ from the contacts is a linear function of the coordinate x: VH ␸͑␰͒ϭ ͩ ␨ͱ1Ϫ␨2ϩ ϩarcsin ␨ͪ ␮ ͒ϭ ϩ⌬ϩ ͒ ␴ ӷ␴ ␲ 2 ͑x eVH͑x a /2, xy xx , ␲ Љ 2͑ Ϫ␨2͒3/2 Ϫaϩ⌬рxрϩaϩ⌬, aϽw ͑21͒ 2 en⌬a 1 ϩ , ␨ϭ␰/a, ͑26͒ 6␬ ͑the parameter ⌬ is defined below͒. Expression ͑21͒ and the definition of ␮(x) for electrons and in a magnetic field give a basis for formulating the problem Ϫ␲2enЈ a2/␬ϭV . ͑27͒ of the properties of incompressible strips in the presence of ⌬ H ͑ ͒ ͑ ͒ transport. In a simplified version of the solution of this prob- Requirements 27 and 24 determine the constant EH ͑ ͒ ϭ ␲ ͑ ͒ ͑ ͒ ͑ ͒ lem it is proposed in analogy with Refs. 8 and 9 to start by 2VH / a in 24 . Formulas 24 – 27 complete the solu- modeling the electron density ␦n(x) inside an incompress- tion of the problem of the behavior of the main channel near ible strip so that its effective value is a constant, correspond- an extremum of the electron density with an even profile ␦ Þ ing to an integer filling factor. Having n(x), we reconstruct n(x) in the presence of VH 0. the distribution ␸(x) with additional requirements that the 2. Recent experiments with Hall samples under condi- tangential electric fields vanish at the boundaries of the strip. tions of the QHE with the use of a scanning potential The ‘‘tie-in’’ of ␸(x) to the behavior of ␮(x) occurs at these microscope13,14 attest to the nonuniform but regular distribu- Þ same ends. In the problem with VH 0 such an algorithm can tion of the equilibrium electron density in them, with a struc- be used only for the central strip. As to the ‘‘side’’ channels, ture of the form n(x) ͑x is the coordinate over a cross section in the transport version the extension of the concepts of Refs. of the sample normal to the current lines͒. The nonuniformity 8 and 9 encounters difficulties that require special discussion. of n(x) is due to contact phenomena which with a high prob- Thus we assume ability are present in the experiments under discussion. In 2 this case for Hall samples 2w2L (Lӷw) far from the con- ␦n͑x͒ϭ͑␯⌬Ϫl͒n ϩnЈ ␰ϩnЉ ␰ /2, H ⌬ ⌬ tacts the distribution n(x) has the form ͑9͒. Using this infor- ␰ϭxϩ⌬, lϭ1,2,3,... ͑22͒ mation, we rewrite a number of general definitions. The quantity ⌬ is given by ␯͑x͒ϭ␲l2 n͑x͒, n ϭ1/͑␲l2 ͒, Ϫaϩ⌬рxрϩaϩ⌬, H H H 3 c eVH w where l is the magnetic length, ⌬ is the overall shift of the ⌬ϭϪ . ͑28͒ H ␲2 W a2 central strip relative to the coordinate origin, and ␯⌬ϵ␯(x ϭ⌬). In the limit ⌬→0 expression ͑22͒ agrees with ␦n(x) In addition, we note that from Ref. 9. ͑xϩ⌬͒ϩa ␦␸ϭ␸ ͒Ϫ␸ ͒Ӎ The corresponding distribution Ex(x) follows from Pois- ͑x,VH ͑x VH , son’s equation written in a form which takes into account the 2 constancy of the electrochemical potential within the limits Ϫaϩ⌬рxрϩaϩ⌬͑29͒ of an integer strip: varies in an approximately linear manner over the cross sec- ␬ ϩa ͒ Ex͑s ␦ ͑␰͒ϭϪ ͵ tion of the central strip between the values 0,VH . e n 2 ds . 2␲ Ϫa ␰Ϫs Turning to the experiments, we shall interpret the most complete Hall data, which are presented in Ref. 14 ͑see Sec. As a result 2 of that paper͒ for the distribution ␦␸(x) over a cross sec- 2␲e ␰ ␰ tion in the central part of a Hall sample of width 2w ͱ͑ Ϫ␰2 2͒ϭ ͭͫ͑␯ Ϫ ͒ ͩ ͪ ϩ Ј ͩ ͪ E 1 /a ⌬ l n T n⌬aT ϭ ␮ x ␬ H 1 a 2 a 10 m in various magnetic fields. The experiment was done using a scanning microscope, which made it possible to 2 n⌬Љ a ␰ ␰ work in the linear region eV рប␻ . However, the tunnel- ϩ ͫT ͩ ͪ ϩT ͩ ͪͬͮ ϩE , H c 8 3 a 1 a H ing technique turns out to be poorly suited for equilibrium measurements ͑as was the case in Refs. 11 and 12͒.Asa ͑ ͒ 23 result, the authors of Ref. 14 published only equilibrium data 16 ␦␸ ͑ ͒ where Ti(s) are Chebyshev polynomials, and the constant for (x), analogous to the difference 29 . These functions EH is as yet arbitrary and will be related to VH below. are reproduced here in Fig. 6. The numbers 1–11 on the The field ͑23͒ can vanish at the ends ␰ϭϮa if right-hand side of Fig. 6 enumerate the recordings of ␦␸(x) 788 Low Temp. Phys. 27 (99–10), September–October 2001 V. B. Shikin

grows. This behavior of a(H) is possible only if the local density n(x) decreases with distance from the center of the sample. In other words, the data on the behavior of a(H)on lines 2–6 are evidence for a version of the contact interaction that leads to depletion of the 2D system. The positions of the boundaries Ϯa(H) with changing field in approximation ͑14͒ are indicated by arrows on lines 3–6. Here the free ប␻ parameter W/ c is chosen such that under conditions cor- responding to the decay of the central strip into two satel- ͑ ͒ lites, the value of amax 15 would be approximately equal ͑as is seen in Fig. 6͒ to one-half of ␻ ͑line 6͒. Hence and from Eq. ͑15͒ it follows that W 3c w 3 24c ϭ ϭ ͑ ͒ ប␻ ␲ ͩ ͪ ␲ . 30 c amax The variation of ␦␸(x) observed in Ref. 14 lies roughly in the region between the arrows ͑see Fig. 6͒. As follows from Eq. ͑29͒, this distribution is approximately linear be- tween the ‘‘banks’’ of a strip. The small asymmetry is due to the influence of the Hall voltage on the boundaries of the central incompressible strip. As an estimate of the asymme- try we use the quantity ⌬͑28͒. With allowance for ͑30͒, the scale of the cyclotron energy on line 5 and the experimental Ӎ рប␻(5) ͑ ͒ ⌬ value eVH 0.01 eV c see Ref. 14 , we have for ⌬͑5͒р0.1w. ͑31͒ The position of the boundaries of the central strip, corrected by ⌬, are indicated by the double arrows on lines 3–6 in Fig. 6. The current asymmetry on lines 7–11 is particularly no- 14 FIG. 6. Recorded traces of the local values of the electric potential along ticeable. Here the central part has undergone a transition to one of the Hall cross sections in the central region of the sample for various magnetic fields. The numbers on the right enumerate the traces, and those on the ‘‘normal’’ state, i.e., is equipotential. The incompressible the left give the local values of the filling factor ␯(0) at the point of the satellites are asymmetrically squeezed toward the boundaries maximum value of the electron density on the assumption that ␯(0)ϭ2on of the Hall region. Unfortunately, the equilibrium treatment line 2 and that the Hall voltage is sufficiently small. The arrows on lines 2–5 of Ref. 9 does not admit generalization to the case of the side indicate the boundaries of the central incompressible strip in the absence of Þ transport current, as calculated using Eq. ͑12͒. The double arrows on line 6 satellites for VH 0. This circumstance was also mentioned give an idea of the scale of the asymmetry in the position of the boundaries above, and so we show without comment the interesting data under the influence of the transport current. Inset: Hall resistance of the on the splitting of the central channel ͑lines 7–11͒. sample versus the magnetic field in the neighborhood of the investigated quantum plateau.14 Also shown here is a selection of lines 1–11 of the main In summary, we have proposed a formalism that gener- figure, allowing one to judge the relationship between the ‘‘quality’’ of the alizes the results of Refs. 8 and 9 for integer ͑incompress- plateau and the width of the incompressible strips corresponding to these ible͒ strips in nonuniform 2D systems to the case of a finite lines. Þ Hall voltage VH 0. The results of the theory are used for interpretation of the data obtained in Refs. 11, 12, and 14 with the use of the hypothesis that the nonuniformity of the for different values of the magnetic field. These same num- Hall sample is of contact origin. This hypothesis is very pro- bers are indicated in the inset, which gives an idea of the ductive for discussing the equilibrium properties of nonuni- ͑ ͒ behavior of Rxy of this sample in the neighborhood of the form Corbino disks with integer strips see Figs. 3 and 4 . Hall plateau. The appearance of the same channels in Hall samples is com- The numbers in the left-hand part of Fig. 6 give the local pletely natural. values of the filling factor ␯(0) at each ith line. These num- The picture that appears is qualitatively quite similar to bers differ from those given in Ref. 14 by a linear shift and the observations and explains many of the existing details. In have been chosen from considerations of self-consistency of particular, we have determined the sign of the contact energy the contact scenario. In our version of events the nucleation for the samples from Ref. 14, which corresponds to depletion of a quasi-symmetric incompressible strip occurs in the of the 2D region, and its scale, which is much larger than the neighborhood of line 2 ͑line 1 does not show any signs of the cyclotron energy ͓see the definition of w in ͑30͔͒. We have existence of a strip, and on line 3 the strip already has a finite given an explanation for the shift ⌬ of the central strip under ͒ ␯ Þ ͓ ͑ ͒ width . Consequently, on this line the value of (0) should the influence of a nonzero Hall voltage VH 0 formulas 28 take the value 2, and this value is taken as a basis for nor- and ͑31͔͒ and also for the shift of the extrema of ␸(x) under ␯ Þ malizing the values of (0). With increasing number i the the influence of VH 0, for which a clear realization is not magnetic field decreases, and the width of the central strip yet available. Low Temp. Phys. 27 (99–10), September–October 2001 V. B. Shikin 789

This study was supported in part by the Russian Foun- 8 D. B. Chklovskii, B. I. Shklovskii, and L. I. Glazman, Phys. Rev. B 46, dation for Basic Research, Grant 01 02 16467. 4026 ͑1992͒. 9 D. B. Chklovskii, K. F. Matveev, and B. I. Shklovskii, Phys. Rev. B 47, ͑ ͒ * 12605 1993 . E-mail: [email protected] 10 P. Fonten, P. Hedriks, F. Blom, J. Wolter, L. Giling, and C. Beenakker, Surf. Sci. 263,91͑1992͒. 11 R. Knott, W. Dietsche, K. von Klitzing, K. Ploog, and K. Eberl, Semicond. 1 L. D. Landau and E. M. Lifshitz, Electrodynamics of Continuous Media, Sci. Technol. 10,117͑1995͒. 2nd ed., rev. and enl., by E. M. Lifshitz and L. P. Pitaevskii ͓Pergamon 12 W. Dietsche, K. von Klitzing, and K. Ploog, Surf. Sci. 361,289͑1996͒. Press, Oxford ͑1984͒; Nauka, Moscow ͑1982͔͒. 13 K. McCormick, M. Woodside, M. Huang, M. Wu, P. McEuen, C. Duruoz, 2 S. Petrosyan and A. Shik, Zh. E´ ksp. Teor. Fiz. 96, 2229 ͑1989͓͒Sov. Phys. and J. Harris, Jr. Phys. Rev. B 59, 4654 ͑1999͒. JETP 69,261͑1989͔͒. 14 P. Weitz, E. Ahlswede, J. Weis, K. V. Klitzing, and K. Eberl, Physica E 6, 3 A. Shik, Fiz. Tekh. Poluprovodn. 29, 1345 ͑1995͓͒Semiconductors 29, 247 ͑2000͒. 697 ͑1995͔͒. 15 A. H. MacDonald, T. M. Rice, and W. F. Brinkman, Phys. Rev. B 28, 3648 4 V. Shikin and N. Shikina, JETP Lett. 62,894͑1995͒. ͑1983͒. 5 S. Laux, D. Frank, and F. Stern, Surf. Sci. 196,101͑1988͒. 16 I. S. Gradshteyn and I. M. Ryzhik, Tables of Integrals, Series, and Prod- 6 V. Shikin, Zh. E´ ksp. Teor. Fiz. 96, 1406 ͑1989͓͒Sov. Phys. JETP 69, 797 ucts ͓Academic Press, New York ͑1980͒; Nauka, Moscow ͑1971͔͒. ͑1989͔͒. 7 I. Larkin and V. Shikin, Phys. Lett. A 151, 335 ͑1990͒. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

On the nature of the half-integer quantum features on the transport and Hall resistances of 2D electron systems in a quantizing magnetic field * E. A. Pashitski Institute of Physics, National Academy of Sciences of Ukraine, pr. Nauki 46, 03028 Kiev, Ukraine ͑Submitted June 18, 2001͒ Fiz. Nizk. Temp. 27, 1069–1082 ͑September–October 2001͒ It is shown that in two-dimensional ͑2D͒ electron systems in a quantizing magnetic field, in addition to electron–hole ͑excitonic͒ pairing through the Coulomb interaction it is also possible to have Cooper pairing of the 2D electrons due to their interaction with 2D phonons and 2D plasmons localized at the interface of the crystals in semiconductor heterostructures. By summing the divergent ͑as T→0͒ ladder diagrams in the zero-sound and Cooper channels, the critical temperatures of the transition to the excitonic and Cooper phases are found. It is shown that the excitonic phase can exist only in comparatively narrow regions near half-integer values of the filling factor ␯ϭ(2nϪ1)/2 (nϭ1,2,3,...), a finding which is in qualitative agreement with the experimentally observed anisotropic features in the transport resistance of 2D systems in the integer quantum Hall effect regime. By taking into account a superposition of states of bound electron pairs and unpaired 2D electrons in the Cooper phase, one can describe practically all of the experimentally observed quantum features in the fractional quantum Hall effect regime, including at values of ␯ that are not described by the composite fermion model. At the same time, the interelectron attraction can promote triplet Cooper pairing of composite fermions, which is accompanied by the appearance of an ‘‘exotic’’ quantum feature at ␯ϭ5/2. Arguments supporting the possibility of experimental observation of the Cooper phase are presented. © 2001 American Institute of Physics. ͓DOI: ͓DOI: 10.1063/1.1414567͔

1. INTRODUCTION of Cooper pairing of 2D electrons in a quantizing magnetic field as a result of the electron–phonon interaction was con- Experimental studies of the integer and fractional quan- sidered previously in Ref. 15. It was shown in the self- tum Hall effects ͑IQHE and FQHE͒ in semiconductor consistent field approximation that in a high enough quantiz- 1,2 ␻ Ͼ␻ ͑ heterostructures show evidence that as the quality of the ing magnetic field under the condition He ˜ ph where ␻ ϭ samples is improved and the mobility of the two-dimensional He eH/cme* is the cyclotron frequency of electrons with ͑ ͒ ␻ 2D electrons in the inversion layers is increased, ever larger effective mass me* , and ˜ ph is the maximum frequency of numbers of new quantum features are observed on the curves the phonon spectrum͒, when in the process of the electron– ϭ of the Hall resistance Rxy RH and transport resistance Rxx phonon interaction the electrons remain on the upper par- ͑ ͒ or Ryy as functions of the strength of the quantizing mag- tially filled Landau level, the critical temperature of the tran- netic field H.3,4 In the FQHE regime one often observes sition to the superconducting state realized on the Landau fractional values of the electronic filling factor ␯ϭq/k which wave functions is determined by the expression cannot be described in the framework of the Laughlin N W2 tanh͓1/2 ln͑1/␯ Ϫ1͔͒ theory,5 on the basis of the fractional statistics of anyons,6–8 ϭ L ph s ͑ ͒ Tc ͑ ␯ Ϫ ͒ , 1 nor with the use of the model of composite fermions,9,10 kB ln 1/ s 1 ␯ϭ Ϯ which leads to fractions of the form n/(2n 1) for n where Wph is the matrix element of the electron–phonon ϭ ␯ 2,3,4,... . interaction, s is the degree of filling of the upper Landau Of greatest interest recently is the quantum feature at the р␯ р ϭ 1␲ 2 level s (0 s 1), NL 2 lH is the degeneracy of the ␯ϭ ϭͱប half-integer value 5/2, which was first observed in Ref. Landau level, and lH c/eH is the quantum magnetic ͑ ប 11 for an isolated GaSb–InAs–GaSb quantum well and then length Planck’s constant and Boltzmann’s constant kB will in GaAs/AlGaAs heterostructures with a high electron henceforth be taken equal to unity͒. 12,13 ␯ϭ ͑ ͒ mobility. Formally the value 5/2 belongs to the re- It follows from Eq. 1 that Tc reaches a maximum at ␯ ϭ ␯ → gion of the IQHE, but the state corresponding to it has all the s 1/2 and goes to zero by a logarithmic law for s 0 and ␯ → ͑ ͒ characteristic properties of a state of the FQHE and has an s 1 Fig. 1 . As a result of the change in the number of energy gap in the spectrum of elementary excitations.3 To filled Landau levels nϭsϩ1 with increasing magnetic field, ␯ϭ ⌬ ϭ interpret the ‘‘exotic’’ quantum state with factor 5/2, in Tc is an oscillatory function of H with period H H/s, and Ref. 14 it was proposed to use a model of p-wave triplet with increasing H the maximum values of Tc increase by a ϳ Cooper pairing of composite fermions due to a weak attrac- linear law, since NL H. tion arising as a result of ‘‘overscreening’’ of the Coulomb An important feature of expression ͑1͒ is that in a suffi- repulsion. ciently strong quantizing magnetic field the value of Tc is In this connection it should be noted that the possibility independent of the 2D electron density of states N2D

1063-777X/2001/27(9–10)/10/$20.00790 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 E. A. Pashitski 791

to electron–hole ͑excitonic͒ pairing and to scattering of the current carriers on quantum charge-density waves ͑CDWs͒, which were predicted in Ref. 26 and studied in more detail in Refs. 27–31. In the self-consistent-field approximation the critical temperature TCDW of the transition to the spatially inhomogeneous excitonic phase is given by the expression26

ϭ ˜ ͒␯ Ϫ␯ ͒ ͑ ͒ TCDW 2NLVc͑q s͑1 s , 2 ˜ where Vc(q) is the matrix element of the screened Coulomb interaction in a quantizing magnetic field at a value of the momentum transfer q equal to the inverse period of the ͑ ͒ CDW. It follows that TCDW , like Tc in Eq. 1 , is a periodic ␯ FIG. 1. Dependence on s of the critical temperatures Tc and TCDW nor- function of magnetic field and reaches maximum values at 2 ͑ ͒ ͑ ͒ malized to NLWph , according to formulas 1 and 5 for different relation- ␯ ϭ half-filling of the upper Landau level, when s 1/2. ships between the matrix elements of the electron–phonon and Coulomb ␯ 2 Ͻ˜ However, the parabolic behavior of TCDW( s) in the in- interactions. Curve 1 for TCDW corresponds to the condition Wph Vc , р␯ р ͑ ͒ 2 ϭ˜ 2 Ͼ˜ terval 0 s 1 Fig. 1 does not agree with the character of curve 2 to Wph Vc , and curve 3 to Wph Vc . the experimentally observed features of Rxx and Ryy in com- paratively narrow regions in the neighborhood of the points ␯ϭ(2nϪ1)/2 for nу3.25 This creates some difficulties for ϭ ␲ me*/2 . This means that the Cooper pairing of electrons interpreting these features as being a manifestation of exci- on the same Landau level is possible for arbitrarily small me* tonic pairing and the related formation of quantum if the interelectron attraction due to the electron–phonon in- CDWs.28–31 Moreover, it remains unclear why no features of ␯ϭ teraction exceeds the Coulomb repulsion. We note that the Rxx and Ryy are observed for 3/2 and 1/2. problem of Cooper pairing of 2D electrons in a quantizing In the present paper we make the conjecture that the magnetic field and the related phenomenon of giant oscilla- features of the transport resistance of a 2D system in the tions of the superconducting order parameter as a function of IQHE regime near ␯ϭ5/2, 7/2, 9/2,... and also the ‘‘exotic’’ ␯ϭ H were considered in Refs. 16 and 17. quantum feature of Rxx and RH at 5/2 and a number of In Ref. 18 it was pointed out that in semiconductor het- other features at values of ␯ in the FQHE regime are conse- erostructures the strongest electron–phonon interaction for quences of an effective interelectron attraction which is due 2D electrons is not with the three-dimensional ͑3D͒ bulk to the interaction of 2D electrons with 2D phonons and 2D phonons but with surface acoustic and optical 2D phonons, plasmons localized at the heterojunction. This attraction which should exist at the interfaces between the different leads to a competition between the electron–hole ͑excitonic͒ crystals. In particular, optical surface oscillations ͑interfacial pairing due to the Coulomb interaction and the electron– modes͒ have been observed experimentally in the Raman electron ͑Cooper͒ pairing due to the electron–phonon and scattering spectra of semiconductor superlattices.19,20 It is electron–plasmon interactions, because of the fact that the also possible that the superconducting transition at TϽ1K corresponding diagrams of the thermodynamic perturbation observed21 in a silicon metal–insulator–semiconductor theory for the two-particle Green’s function at zero total en- ͑MIS͒ structure ͑in the absence of magnetic field͒ is the re- ergy and T→0 have divergences of the same power sult of an electron–phonon interaction with surface phonons. (ϳ1/T) in the zero-sound and Cooper channels. Another possible mechanism of attraction between 2D By simultaneous summation of the divergent ladder dia- electrons, which was considered previously in Refs. 22 and grams in both channels in the local-interaction approxima- 23, is the ‘‘dynamic overscreening’’ of the Coulomb interac- tion the critical temperatures of the transition to the excitonic tion in layered semiconductor structures and quantum wells, and Cooper phases are found, and the phase diagrams of the an effect which arises due to the exchange of virtual quanta 2D system are constructed in the IQHE and FQHE regimes. of the low-frequency plasma oscillations of ‘‘heavy’’ holes It is shown that the excitonic phase with a spatially periodic ӷ with an effective mass mh* me* . This electron–plasmon in- quantum CDW can exist only in comparatively narrow re- teraction mechanism is radically different from that proposed gions around the half-integer values ␯ϭ(2nϪ1)/2, in quali- in Ref. 24 for explaining the experimental data21 on the plas- tative agreement with the experimental data25 on the obser- mon mechanism of Cooper pairing of 2D electrons due to vation of the features of Rxx and Ryy in the IQHE regime. their interaction with intrinsic high-frequency 2D plasmons In the ultraquantum limit, by taking into account a su- ␻ ϳͱ with a square-root dispersion relation q q. perposition of states of the bound electron pairs and the un- In addition, in a GaAs/AlGaAs heterostructure with a paired 2D electrons on the lowest spin Landau level, one high mobility of the 2D electrons in the IQHE regime at can, on the basis of the Halperin multiparticle wave ultralow temperatures, new quantum features of the transport function,32 describe practically all of the experimentally ob- resistance have been observed near half-integer values of the served quantum features of the Hall and transport resis- electronic filling factor ␯ϭ(2nϪ1)/2, starting with nϭ3.25 tances, including those at values of the filling factor ␯ Such features, in the form of alternating maxima and minima ϭ2/7, 2/9, 2/11, 3/11, etc., which are not described by the 9,10 or narrow peaks of Rxx(H) and Ryy(H) in the region of the composite fermion model. At the same time, the interelec- broad plateau on the field dependence of RH , are character- tron attraction due to the electron–phonon and electron– Þ ized by strong anisotropy (Ryy Rxx) and are apparently due plasmon interactions with surface phonons and plasmons can 792 Low Temp. Phys. 27 (99–10), September–October 2001 E. A. Pashitski bring about triplet Cooper pairing of composite fermions, which is accompanied by the appearance of an ‘‘exotic’’ quantum feature at ␯ϭ5/2.11–14

2. COULOMB, ELECTRON–PHONON AND ELECTRON–PLASMON INTERACTIONS IN 2D SYSTEMS IN A QUANTIZING MAGNETIC FIELD

In studying the properties of 2D electron systems in a quantizing magnetic field, one ordinarily takes into account only the Coulomb interaction between electrons26–31 with a 2 ␧ ͑ ␧ characteristic energy e / 0lH where 0 is the dielectric con- FIG. 2. Antisymmetrized matrix element of the Coulomb interaction, stant of the crystal͒. ˜ ␲ 2 ␧ Us(q), normalized by e lH / 0 , as a function of the dimensionless mo- As was shown in Ref. 33, the Hamiltonian of the ϭ ϭ ͑ ͒ ϭ ͑ ͒ mentum x qlH for s 0 dashed curve and s 1 solid curve . Coulomb interaction between 2D electrons found on the same Landau level in the momentum representation has a certain symmetry and is invariant with respect to a certain 33 identity integral transformation, which enables one to elimi- the Coulomb interaction. An analogous sign change of ˜ у Ϫ1 nate all loop diagrams from the thermodynamic perturbation Us(q) for q lH should occur on all Landau levels. ͑ series with the aid of the operation of antisymmetrization of As we have said, in real heterostructures p – n junc- ͒ the Fourier components of the Coulomb potential. tions in addition to the Coulomb interaction one must take The antisymmitrized potential for the sth Landau level into account the inelastic interaction of 2D electrons in the has the following form:33 inversion layers with surface phonons localized at the inter- faces of the crystals18–20 and also with surface plasmons, 1 U˜ ͑q͒ϭ ͓˜V ͑q͒Ϫ˜VЈ͑q͔͒, ͑3͒ which are associated with the relatively low-frequency col- s 2 s s lective oscillations of the ‘‘heavy’’ holes.22,23 In particular, as where was shown in Ref. 22, at the interface between two semicon- ductors, one p-type and the other n-type, with substantially ˜ ͑ ͒ϭ ͑ ͒ 2͑ ͒ Vs q Vc q ws q ; different effective masses of the holes and electrons and with a high mobility of free carriers in a certain region of frequen- 2 lH cies ␻ and longitudinal wave numbers q, there can exist ˜VЈ͑q͒ϭ ͵ d2p exp͕ip ql2 ͖˜V ͑p͒; ͑4͒ s 2␲ • H s weakly damped surface plasmons with the following disper- → Ϫ 2 2 sion relation, which is of the acoustic type at q 0: ϭ q lH/4 2 2 ws(q) e Ls(q lH), Ls(z) is a Laguerre polynomial, ϭ ␲ 2 ␧ ␻ ͑ ͒ϭ⍀ ͓ ϩͱ ϩ 2 2͔Ϫ1 ͑ ͒ and Vc(q) 2 e /q 0 is the Fourier component of the un- pl q h 1 1 1/q Re , 7 screened Coulomb repulsion in the 2D system. In particular, where ⍀ ϭͱ4␲e2n /␧ m* is the plasma frequency of for the lowest Landau level (sϭ0) in the ultraquantum limit h h 0 h ͑ ͒ ͑ ͒ ‘‘heavy’’ holes with an average volume concentration nh , the potential 3 with allowance for 4 is expressed as ϭͱ ␧ ␲ 2 and Re EF 0/6 e ne is the screening radius in the case of ␲ 2 Ϫx2/2 screening by degenerate ‘‘light’’ electrons with a concentra- e lH e 2 U˜ ͑q͒ϭ ͫ Ϫͱ␲/2 I ͑x2/4͒eϪx /4ͬ, ͑5͒ 0 ␧ x 0 tion ne and Fermi energy EF in crystals with the same 0 ␧ dielectric constant 0 . Here the electric field of the ϭ where x qlH , and I0(z) is a modified Bessel function of the surface plasma wave falls off with increasing distance z from first kind. In the case of two filled Landau levels (sϭ1) we the interface according to an exponential law have, according to ͑3͒ and ͑4͒, ϳ Ϫ ͱ 2ϩ Ϫ2 exp( z q Re ), and the Hamiltonian of the adiabatic Ϫ 2 ͑under the condition m*ӷm*) electron–plasmon interaction ␲e2l e x /2 x2 h e ˜ ͑ ͒ϭ H ͭ ͑ Ϫ 2͒2Ϫͱ␲ ͫ Ϫx2/4 is analogous to the Fro¨lich Hamiltonian for the electron– U1 q ␧ 1 x /2 I0 e 0 x 4 phonon interaction: 3 x2 5 x2 Ϫ ⌽ͩ Ϫ ͪ ϩ ⌽ͩ Ϫ ͪͬͮ ͑ ͒ ͑ ͒ϭ ͑ ͒␻1/2͑ ͒ ϩ ϩϩ ͑ ͒ 2 ,1; 6 ,1; , 6 Hpl z i ͚ gpl q,z pl q akϩq,␴ak␴bq h.c., 8 2 2 2 2 k,q,␴ ⌽ ␣ ␤ ϵ ␣ ␤ where ( , ;z) 1F1( , ;z) is the confluent hypergeo- where metric function. ␲ 2 1/2 ˜ e Ϫ Figure 2 shows the momentum dependence of U0(q) g ͑q,z͒ϭͩ ͪ exp͑Ϫzͱq2ϩR 2͒, ͑9͒ pl Ϫ e ˜ ˜ Ͼ Ͻ ␧ ͱq2ϩR 2 and U1(q). We see that U0(q) 0 in the region q q0 0 e Ϸ Ϫ1 ˜ Ͼ Ͻ Ϸ Ϫ1 0.7lH , and U1(q) 0 in the region q q1 1.4lH , and z is the coordinate of the 2D electron layer ͑see below͒. which corresponds to Coulomb repulsion. In the regions The corresponding effective Hamiltonian of the Ͼ Ͼ ˜ ˜ q q0 and q q1 , however, the potentials U0(q) and U1(q) electron–electron interaction in second order of perturbation become negative, corresponding to an effective attraction be- theory owing to the exchange of virtual 2D phonons and 2D tween 2D electrons due to the ‘‘magnetic overscreening’’ of plasmons in the Landau gauge has the form ͑cf. Ref. 33͒ Low Temp. Phys. 27 (99–10), September–October 2001 E. A. Pashitski 793

In this case, if the width of the localization region for ϭ ͓ 2 ͑ ͒ ͑ ␧ Ϫ␧ ͒ Hee ͚ ͚ ˜gph q Dph q, s sЈ ‘‘heavy’’ holes satisfies L ӷL , the frequency of the adia- Ј Ј Ј h e k,k ,q s,s batic surface plasmons is given by ϩ 2 ͒d ␧ Ϫ␧ ͒ ˜g ͑q Dpl͑q, s sЈ ͔ pl ␻ ͑ ͒ϭ⍀ ͱ ͑ ϩ ͒ ͑ ͒ pl q h / 2 1 1/qae* , 14 ϫ ͒ ͒ ϩ ϩ ͑ ͒ ws͑q wsЈ͑q a ϩ a Ϫ akЈsЈaks , 10 k q,s k q,s ␻ ϳͱ → so that s(q) q for q 0, and the electron–plasmon in- where ˜gph and ˜gpl are the matrix elements of the electron– teraction constant has the form ϩ phonon and electron–plasmon interactions, a and aks are ks 2 1/2 the creation and annihilation operators for an electron with 2␲e 2 ˜g ͑q͒ϭͩ ͪ expͭ Ϫ˜dͩ qϩ ͪͮ , ͑15͒ pl ␧ ͑ ϩ ͒ 2D momentum k at the sth Landau level, and Dph and Dpl 0 q 2/ae* ae* are the phonon and plasmon Green’s functions: where ˜d is the distance between the maxima of the density 2 ␻ ͑q͒ distributions for electrons and holes localized on opposite ͑ ␻͒ϭ ph,pl ͑ ͒ Dph,pl q, ␻2Ϫ␻2 ͑ ͒ . 11 ˜ Ͼ ph,pl q sides of the heterojunction (d d). In a magnetic field strong enough to quantize the spec- ␻ If the characteristic energies of the surface phonons ph(q) trum of ‘‘light’’ 2D electrons but not the ‘‘heavy’’ holes, the ␻ ϳ Ϫ1 and plasmons pl(q) for q lH do not exceed a value spectrum of 2D plasmons acquires a finite gap at q→0: ␮ ͑ ␮ ͒ 2 BH where B is the Bohr magneton , then the inelastic ␻ ͑ ͒ϭͱ␻2 ϩ␻2 ͑ ͒ ͑ ͒ processes due to the electron–phonon and electron–plasmon ˜ pl q Hh pl q , 16 interactions do not lead to mixing of the states of different ␻ ϭ Landau levels. In that case we may neglect retardation ef- where Hh eH/mh*c is the cyclotron frequency of the fects in ͑11͒ for 2D electrons on the same Landau level (s holes. We note that the surface plasmon energy ͑16͒ for ϭ ␻ϭ ϭ ϭϪ m*Ͻm exceeds the value of the Zeeman splitting of the sЈ) and set 0, so that Dph(q,0) Dpl(q,0) 1, corre- h 0 sponding to interelectron attraction. Landau levels, so that the inelastic electron–plasmon inter- The effective attraction between 2D electrons due to action can lead to mixing of the states of the nearest spin their interaction with surface phonons is characterized by the Landau levels. square of the matrix element of the electron–phonon inter- Since the screening of the Coulomb interaction of the action, which in the long-wavelength approximation can be electrons is strongly suppressed in a quantizing magnetic ␻ Ϸ⍀ & written in the form field, the frequency of the 2D plasmons is pl h / , and the square of the matrix element of the electron–phonon in- 2 ˜ 2 ␥2 teraction has the form ˜a DA ˜ 0 ˜g2 ͑q͒ϭ ͫ ϩ ͬ eϪ2qd, ͑12͒ ph 2 2 2 Ms˜ ˜ ␻2 2␲e ˜ M ˜ 0 2 ͑ ͒ϭ Ϫ2qd ͑ ͒ ˜gpl q ␧ e . 17 q 0 where ˜a is the average lattice constant at the heterojunction, d is the distance of the 2D electron layer from the interface Taking into account that the Hamiltonian ͑10͒ for s ˜ ϭsЈ has the same symmetry as the Coulomb Hamiltonian,33 of the crystals, DA is the deformation potential of acoustic ␥ by analogy with the Coulomb potential ͑6͒ we obtain the 2D phonons with phase velocity ˜s along the interface, ˜ 0 is the deformation constant for optical 2D phonons with fre- following effective antisymmetrized electron–phonon and ␻ ˜ electron–plasmon interaction potential: quency ˜ 0 , and M and M are the total and reduced masses of the atoms in the unit cells of the crystals on the two sides W˜ ͑q͒ХW˜ ph͑q͒ϩW˜ pl͑q͒ of the interface. If the difference of the elastic constants of s s s the crystals is sufficiently small, so that the values of ˜s and 1 l2 ␻ ͑ ͒ ϭϪ ͭ ͓ 2 ͑ ͒ϩ 2 ͑ ͔͒ 2͑ ͒Ϫ H ˜ 0 are small compared with the bulk values , the value of ˜gph q ˜gpl q ws q ␲ 2 2 2 ˜gph can be considerably greater than the electron–phonon interaction with 3D phonons in semiconductors.34 ϫ ͵ 2 ͑ 2 ͓͒ 2 ͑ ͒ϩ 2 ͑ ͔͒ 2͑ ͒ͮ To calculate the matrix element of the electron–plasmon d p exp ip•qlH ˜gph p ˜gpl p ws p . interaction with surface plasmons in heterostructures with ͑ ͒ size quantization of the electron spectrum in the inversion 18 layer, we take into account that for a sufficiently thin electron For calculating the matrix element W˜ (q) it is necessary layer of thickness L the screening radius in the case of a s e to know the explicit momentum dependence of the electron– quadratic spectrum of degenerate 2D electrons with concen- phonon and electron–plasmon interaction constants, since ϭͱ ␲ ˜ tration N and Fermi momentum kF 2 N is given by Re for a contact ͑point͒ interaction with a constant Fourier com- ϭͱ ͑ ϭ␧ 2 Leae*/2 where ae* 0 /me*e is the effective Bohr ra- ponent, expression ͑18͒ is identically zero ͑see Ref. 33͒. ͒ dius of the electron . Here the Fourier component of the In the long-wavelength approximation (qdϽ1) for the Ӷ screened Coulomb potential under the condition Le ae* to electron–phonon interaction constant for surface phonons ͑ ͒ good accuracy reduces to the expression see Ref. 35 one can to good accuracy keep only the correction linear in q ͓see Eq. ͑12͔͒: 2␲e2 2␲e2 ˜ ͑ ͒Х ϭ ͑ ͒ Vc q ϩ 2 ϩ . 13 2 ͑ ͒Х 2 ͑ ͒͑ Ϫ ͒ ͑ ͒ q Le/2Re q 2/ae* ˜gph q ˜gph 0 1 2qd . 19 794 Low Temp. Phys. 27 (99–10), September–October 2001 E. A. Pashitski

4␲e2˜d ͑ ͒Ϫ 2 ͑ ͒ϭ ͑ ͒͑ Ϫ Ϫqd˜͒Ϸ ͑ Ϫ ˜ ͒ Vc q ˜gpl q Vc q 1 e ␧ 1 qd . 0 ͑22͒ As a result of the antisymmetrization the constant term in ͑22͒ drops out, and for the total matrix element of the Coulomb and electron–plasmon interactions, according to Eqs. ͑3͒ and ͑18͒, for sϭ0 and 1 we obtain the same mo- mentum dependence as for the matrix elements of the electron–phonon interaction, Eqs. ͑20͒ and ͑21͒, but with the opposite sign, since the Coulomb repulsion is dominant over FIG. 3. Antisymmetrized matrix element of the electron–phonon interaction the attraction coming from the electron–plasmon interaction. ˜ ph 2 ϭ ϭ Ws (q), normalized to ˜gph(0)d/lH and plotted versus x qlH , for s 0 Here the corresponding coupling constant with allowance for ͑dashed curve͒ and sϭ1 ͑solid curve͒. ͓ 2 Ϫ ␲ 2˜ 2 the electron–phonon interaction is ˜gph(0) 2 e d / ␧ 0d͔d/lH . Figure 4 shows the momentum dependence of the total ˜ As a result, the phonon part of the matrix element Ws(q) potential of the Coulomb, electron–plasmon, and electron– for sϭ0 becomes phonon interactions for sϭ0 ͑a͒ and sϭ1 ͑b͒ for different ␤ϭ 2 ␧ values of the dimensionless parameter ˜gph(0) 0d/ d 2 ␲ 2˜ 2 ˜ ϩ ˜ W˜ ph͑q͒ϭ˜g2 ͑0͒ ͫxeϪx /2Ϫͱ␲/2⌽ 2 e d . We see that the potential Us(q) Ws(q) corre- s ph ͑ ͒ Ͻ lH sponds to attraction is negative in the region qlH 1, with a ϭ ␤Ͼ 2 minimum at the point q 0, if 1, or in the region qlH 1 x 2 ϫͩ Ϫ ,1; ͪ eϪx /2ͬ. ͑20͒ Ͼ1if␤Ͻ1. 2 2 If the adiabatic electron–plasmon interaction with sur- Here we have used the Kummer formula ⌽(␣,␤;z) face plasma oscillations of ‘‘heavy’’ holes is suppressed for ϭez⌽(␤Ϫ␣,␤;Ϫz). For sϭ1 the corresponding matrix el- some reason, e.g., because of a large distance ˜d, in which ement of the electron–phonon interaction is case the electron–plasmon interaction is exponentially small ͓see Eq. ͑17͔͒, then the resulting momentum dependence of d 2 the matrix element of the Coulomb and electron–phonon in- ˜ ph͑ ͒ϭ 2 ͑ ͒ ͭ ͑ Ϫ 2͒2 Ϫx /2Ϫͱ ␲ͫ⌽ W1 q ˜gph 0 x 1 x e 2 ͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ lH teractions, according to Eqs. 5 , 6 , 20 , and 21 , has the form shown in Fig. 5 for sϭ0 ͑a͒ and sϭ1 ͑b͒ for different 2 2 1 x 3 x ˜␤ϭ 2 ␧ ϫͩ Ϫ ,1; ͪ Ϫ6⌽ͩ Ϫ ,1; ͪ ϩ15⌽ values of the dimensionless parameter ˜gph(0) 0d/ 2 2 2 2 ␲ 2 2 e lH characterizing the relative value of the electron– 2 phonon interaction constant and the Coulomb repulsion. We 5 x 2 ϫͩ Ϫ ,1; ͪͬeϪx /2ͮ . ͑21͒ see that in this case there is a certain region of momentum 2 2 ˜ ϩ ˜ ph Ͻ transfers in which attraction exists, Us(q) Ws (q) 0, but ˜ ph for q→0 the Coulomb repulsion is dominant. Figure 3 shows the momentum dependence of W0 (q) ˜ ph Ͻ and W1 (q). We see that in the region qlH 1 the matrix element of the electron–phonon interaction corresponds to Ͼ 3. COMPETITION BETWEEN THE COOPER AND EXCITONIC attraction, while repulsion is dominant for qlH 1. ͑ ͒ MECHANISMS OF PAIRING IN 2D SYSTEMS IN A Using expression 17 , which in fact describes a weak- QUANTIZING MAGNETIC FIELD ening of the Coulomb repulsion by the effective interelectron attraction due to the electron–plasmon interaction, we obtain As will be shown below, the interelectron attraction due in the long-wavelength approximation to the electron–phonon and electron–plasmon interactions,

␲ 2 ␧ FIG. 4. Momentum dependence of the total potential of the Coulomb, electron–plasmon, and electron–phonon interactions, normalized to e lH / 0 ,for ϭ ͑ ͒ ϭ ͑ ͒ ␤ϭ 2 ␧ ␲ 2 ͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ s 0 a and s 1 b and various values of the parameter ˜gph(0) 0/2 e d:0 1 , 0.5 2 , 1.5 3 ,2 4 . Low Temp. Phys. 27 (99–10), September–October 2001 E. A. Pashitski 795

␲ 2 ␧ ϭ ͑ ͒ ϭ ͑ ͒ FIG. 5. Momentum dependence of the total potential of the Coulomb and electron–phonon interactions, normalized to e lH/ 0 ,fors 0 a and s 1 b ˜␤ϭ ␤˜ 2 2 ͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ and various values of the parameter 2 d /lH :0 1 , 0.5 2 , 1.5 3 ,2 4 .

which was discussed in the preceding Section, promotes the Summation of the ladder diagrams of the type ͑25͒ gives a Cooper pairing of 2D electrons on the same spin Landau pole in the two-particle Green’s function at a certain tem- level, and this competes with the excitonic pairing brought perature Tc determined by the relation about by the Coulomb attraction between electrons and tanh͓1/2 ln͑1/␯ Ϫ1͔͒ holes. ϩ ͓ ˜ ͑ ͒ϩ ˜ ͑ ͔͒ s ϭ ͑ ͒ 1 NL Ws q Us q ␯ Ϫ ͒ 0. 26 An analysis of the diagrams for the two-particle Green’s 2Tc ln͑1/ s 1 function of 2D electrons on the sth Landau level at TÞ0 has shown36 that the summation of the ladder diagrams of the An analogous result was obtained in Ref. 15 in the self- form consistent-field approximation ͓cf. Eq. ͑1͔͒. However, since diagrams of the types ͑23͒ and ͑25͒ are ␯ ͑1Ϫ␯ ͒ ␻͒ ␻ϩ␻ Ϫ␻ ͒ϭϪ␦ s s ͑ ͒ characterized by the same divergence ϳ1/T at T→0, to cal- T͚ Gs͑ Gs͑ 1 2 ␻ ␻ 23 ␻ 1 2 T culate the critical temperature one must do a simultaneous summation of these diagrams. The situation here is analo- in the zero-sound channel, which have a power-law diver- gous to the so-called ‘‘parquet’’ approximation for one- gence for T→0 ͑here G (␻) is the single-particle Green’s s dimensional systems, when all of the logarithmically diverg- function and ␦␻ ␻ is the Kronecker delta with respect to the 1 2 ing diagrams are summed ͑see Refs. 37–40͒. ␻ ␻ ͒ discrete Matsubara frequencies 1 and 2 , gives rise to a The integral equation for the vertex part ͑four-pole͒⌫ ͑ ͒ s singularity a simple pole at a certain finite temperature, of the electron–electron interaction at the sth Landau level, which corresponds to the critical temperature TCDW of the obtained with a summation of the divergent zero-sound and ͓ transition to the excitonic phase with a quantum CDW see Cooper diagrams, is presented in graphical form in Fig. 6. Eq. ͑2͔͒. ⌫(0) ͑ → ͒ ⌫ ⌫(1) Here s is the nonsingular at T 0 part of s , and s At the same time, it was stated in Ref. 36 that Cooper (2) and ⌫ are the irreducible parts of ⌫ corresponding to the pairing cannot occur in 2D systems, since the simplest cor- s s Cooper and zero-sound channels. Adopting for simplicity the responding diagram has the form local-interaction approximation, we arrive at an expression 2f͑␧͒Ϫ1 for ⌫ containing a pole with respect to T: ␻͒ ␻ ϩ␻ Ϫ␻͒ϭ ͑ ͒ s T ͚ Gs͑ Gs͑ 1 2 , 24 ␻ i͑␻ ϩ␻ ͒Ϫ2␧ 1 2 ⌫͑0͒ s ␧ ϭ ␧/Tϩ Ϫ1 ⌫ ͑T͒ϭ , ͑27͒ where f ( ) (e 1) is the Fermi distribution function s 1Ϫ␭ F ͑␯ ͒/T Ӷ␮ s s s of the electrons, which under the condition T BH is to ␯ exponential accuracy equal to the degree of filling s of the where ͑ ͒ ␻ ϩ␻ Þ upper Landau level see Refs. 15 and 36 .If 1 2 0, ͑ ͒ → ␭ ϭϪ ˜ ͒ϩ ˜ ͒ ͑ ͒ then expression 24 is nonsingular for T 0 and goes to s 2NL͓Ws͑qs Us͑qs ͔; 28 ␧ ϭ␯ zero at half-filling of the sth Landau level, when f ( ) s ϭ1/2. ␻ ϩ␻ ϭ However, for zero total energy 1 2 0 expression ͑24͒ at the point ␧ϭ0 has an indeterminacy of the type 0/0, ␧ ϭ␯ which can be resolved by using the relations f ( ) s and ␧ϭ ␯ Ϫ Ͻ␯ Ͻ Tln(1/ s 1) in the region 0 s 1. In that case expres- ͑ ͒ ␻ ϭϪ␻ sion 24 under the condition 1 2 takes the form tanh͓1/2 ln͑1/␯ Ϫ1͔͒ ͑␻͒ ͑Ϫ␻͒ϭ s ͑ ͒ T͚ Gs Gs ␯ Ϫ ͒ . 25 ␻ 2T ln͑1/ s 1 ͑ ͒ FIG. 6. Graphical form of the integral equation for the vertex part of the We see that, like the zero-sound diagram 23 , the Coo- interelectron interaction, obtained by summation of the divergent ͑at T→0͒ per diagram ͑25͒ diverges in a power-law manner at T→0. ladder diagrams in the zero-sound and Cooper channels. 796 Low Temp. Phys. 27 (99–10), September–October 2001 E. A. Pashitski

␯ FIG. 7. The function Fs( s) for different relationships between the param- ␥(1) ␥(2) ␥(1)Ͻ␥(2) ͑ ͒ ␥(1)ϭ␥(2) ͑ ͒ ␥(1)Ͼ␥(2) ͑ ͒ eters s and s : s s 1 ; s s 2 ; s s 3 . ␯ (s) (s) FIG. 8. Dependence on s of the critical temperatures Tc and TCDW , ␭ ␥(1) ␥(1)Ͻ␥(2) ␭ normalized to s s , for s s . The value of s for the Cooper phase tanh͓1/2 ln͑1/␯ Ϫ1͔͒ was chosen at the point of maximum interelectron attraction (qϭ0) due to ␯ ͒ϭ␥͑1͒ s Ϫ␥͑2͒␯ Ϫ␯ ͒ the electron–phonon and electron–plasmon interactions, and for the exci- Fs͑ s s͑1 s . s 2ln͑1/␯ Ϫ1͒ s maxϷ Ϫ1 ϭ s tonic phase, at the point of maximum repulsion (q1 1.7lH )fors 1 and ͑29͒ ␤ϭ2 ͑Fig. 4b͒. ␥(1) ␥(2) Here s and s are the relative contributions of the dia- grams of the Cooper and zero-sound types, and qs is the ␯ tem in a quantizing magnetic field in the variables T and s momentum transfer corresponding to the maximum values of р␯ р ␥(1)Ͻ␥(2) ␭ in the interval 0 s 1 under the condition s s .We the coupling constant s and critical temperature: (s) see that the critical temperature of the excitonic phase TCDW ͑␯ ͒ϭ␭ ͑␯ ͒ ͑ ͒ ␯ ϭ Ͻ (s) Ts s sFs s . 30 is maximum at the point s 1/2, so that for T TCDW max in Figure 7 shows curves of the function F (␯ ) con- relatively narrow intervals of magnetic field field H near s s ␯ϭ Ϫ structed according to Eq. ͑29͒ for several different relation- half-integer values of the total filling factor (2n 1)/2 ␥(1) ␥(2) one may observe features in Rxx and Ryy due to the scatter- ships between the parameters s and s . We see that 25,29–31 ␥(1)у␥(2) ␯ ing of electrons on quantum CDWs, and the width of under the condition s s the function Fs( s) is always ␯ ϭ these field intervals should increase with decreasing T. positive and has a minimum at the point s 1/2. In this case The fact that the features of Rxx and Ryy are not ob- Ts is equal to the critical temperature of the transition to the (s) served upon the filling of the two lowest Landau levels, Cooper phase Tc and is nonzero only for a positive value of ␭ Ͼ when ␯ϭ1/2, 3/2 ͑Ref. 25͒, may be due to an insufficiently the coupling constant s 0, which corresponds to the maxi- mum electron–electron attraction, i.e., a negative minimum low measurement temperature or to the circumstance that ␥(1)Ͼ␥(2) in these cases. Here T(s) ϭ0, and T(s)Ͼ0 in the of the potential W˜ (q)ϩU˜ (q) at a momentum transfer q s s CDW c s s entire interval 0р␯ р1, i.e., a transition can occur only to ϭqmin ͑see Figs. 4 and 5͒. Because of the interelectron at- s s the Cooper phase as T is lowered. traction due to the electron–phonon and electron–plasmon It should be noted that the coherent state of Cooper pairs interactions, in the temperature region 0ϽTϽT ϵT(s) there s c in a quantizing magnetic field, analogous to the Cooper con- exists a Cooper phase with a nonzero concentration of bound densate in superconductors,41 can arise only in the case of electron pairs and with a finite energy ͑Cooper͒ gap in the rather strong overlap of their wave functions, i.e., under the spectrum of elementary excitations. We note that when the condition that the average distance between pairs, ¯r electron–plasmon interaction is taken into account, the maxi- p у(2/␲N)1/2, is less than the localization length 2l of the mum attraction is attained for qϭ0 ͑Fig. 4͒, which corre- H electron wave functions. This corresponds to values of the sponds to a spatially homogeneous Cooper phase. In this total filling factor ␯ϭN/N ϭ4l2 /¯r2Ͼ1. case, in spite of the fact that the surface plasmon energy L H p ␻ ˜ pl(q) can exceed the Zeeman splitting of the Landau levels ͓ ͑ ͔͒ (s)Ӷ ␮ 4. COOPER PAIRING OF 2D ELECTRONS AND THE PHASE see Eq. 16 , under the condition Tc 2 BH the Cooper pairing of 2D electrons is possible only for electrons on the DIAGRAM IN THE ULTRAQUANTUM LIMIT same partially filled spin Landau level and is therefore of a A question of particular interest is that of the Cooper triplet p-wave character. pairing of 2D electrons in the ultraquantum limit (sϭ0), ␥(1)Ͻ␥(2) Under the condition s s in the neighborhood of when the filling factor ␯ of the lowest spin Landau level is ␯ ϭ ␯ ͑ 0 the point s 1/2 the function Fs( s) becomes negative Fig. ␯ϭ ␲ 2 Ͻ equal to 2 lHN 1, so that the average distance between ͒ ␯ Ͻ Ϫ 7 . In the region of magnetic fields where Fs( s) 0 there electrons ¯r ϭ(␲N) 1/2Ͼl ͱ2. ␭ Ͻ e H can be nonzero values of Ts only for s 0, which corre- In this case the phase diagram of the 2D system can be ˜ ␯ϭ sponds to a positive maximum of the potential Ws(q) presented in the variables T – H, since H0 /H, where ϩ ˜ ϭ max ͑ ͒ ϭ ⌽ ⌽ Us(q) at the point q qs see Figs. 4 and 5 , i.e., H0 N 0 and 0 is the magnetic flux quantum. Figure 9 electron–electron repulsion and electron–hole attraction. In shows the phase diagrams corresponding to the conditions (s) ␥(1)Ͼ␥(2) ͑ ͒ ␥(1)Ͻ␥(2) ͑ ͒ this case Ts is equal to the critical temperature TCDW of the 0 0 a or 0 0 b for the same parameter val- (0) transition to a spatially homogeneous excitonic phase with a ues as in Fig. 8. The temperature region below the Tc (H) quantum CDW. curves is the existence region of the Cooper phase, and be- (0) Figure 8 shows the phase diagram of a 2D electron sys- low the curve TCDW(H) in Fig. 9b is the existence region of Low Temp. Phys. 27 (99–10), September–October 2001 E. A. Pashitski 797

␭ ␥(1) ϭ ␥(1) FIG. 9. Phase diagram of a 2D system in the variables T and H, normalized to 0 0 and H0 , respectively, in the ultraquantum limit (s 0) for 0 у␥(2) ␥(1)Ͻ␥(2) ͑ ͒ ␤ϭ ͑ ͒ 0 and 0 0 b for 2 Fig. 4a .

41 ͑ ͒ ϭ ϭ ϭ ␯ ϭ␯ ϭ the excitonic phase. By analogy with superconductors it is according to 31 , for m 7, p 16, r 2( e p 1/3), the natural to suppose that in the Cooper phase in the region 0 fraction ␯ϭ1/5 for mϭ7, pϭ22, rϭ8 ͑or rϭ12͒ or for m Ͻ Ͻ (0) ϭ ϭ ϭ ␯ ϭ ␯ ϭ ␯ϭ T Tc there is coexistence of bound electron pairs and 7, p 28, r 6( e 2 p 1/2), and the fraction 1/7 → ϭ ϭ ϭ ͑␯ ϭ ␯ ϭ ͒ unbound 2D electrons, whereas for T 0 all of the electrons for m 9, p 46, r 8 e 3/5, p 1/5 . у (0) ϭ ͑ Þ ͒ ͑ ͒ are bound into Cooper pairs, and for T Tc all of the elec- On the other hand, for p 2r but r 2m relation 31 trons are unpaired and form an ‘‘incompressible’’ Laughlin for any m gives a sequence of fractional values ␯ϭ2/r, 2D liquid.5 which for odd values of rϭ(2nϩ1) with nу1 corresponds The possibility of coexistence of bound pairs and un- to the experimentally observed sequence of fractions ␯ bound electrons in a 2D system in the ultraquantum limit ϭ2/3, 2/5, 2/7, etc., while for even rϭ2(2nϩ1) it reduces was considered in Ref. 32 without specifying the pairing to the Laughlin fractions ␯ϭ1/(2nϩ1). We note that the mechanism. With the use of a many-particle variational wave fractions ␯ϭ2/7, 2/9, 2/11, etc. are not described by the function of the polynomial type,5,27 symmetric with respect composite fermion model,9 which gives ␯ϭn/(2nϮ1) for to interchanges of pairs ͑bosons͒ and antisymmetric with re- nу2. spect to the interchanges of free electrons ͑fermions͒, it was The condition pϭ2r, according to Eq. ͑32͒, corresponds ͑␯ shown in Ref. 32 that a superposition of the states N p of the to a state with zero concentration of free 2D electrons e ϭ ␯ ϭ ͒ ϭ bound pairs and Ne of the unbound 2D electrons should lead 0, p 1/2 , i.e., to the Cooper phase at T 0, when all the to quantum features of the Hall resistance RH and transport electrons on the lowest spin Landau level are bound into resistance Rxx in the FQHE regime for values of the filling triplet pairs with spin 1. However, fractions of the type ␯ ␯ϭ ͑ ϭ ϩ ϭ ϩ ͑ ͒ ␯ Þ factor N/N␸ where N Ne 2N p is the total concentra- 2/(2n 1) are contained in relation 31 for e 0 as well. ϭ ⌽ tion of 2D electrons, and N␸ H/ 0 is the number of flux For example, the fraction ␯ϭ2/5 is realized for mϭ7, p ⌽ ͒ ϭ ϭ ␯ ϭ␯ ϭ ␯ϭ ϭ quanta 0 per unit area given by the following relation: 12, r 2( e p 1/3), the fraction 2/7 for m 3, p ϭ ϭ ␯ ϭ ␯ ϭ ϭ ϭ ϭ ͑␯ 4mϩpϪ4r 6, r 9( e 2/3, p 1/6) or for m 7, p 20, r 8 e ␯ϭ ͑ ͒ ϭ ␯ ϭ ͒ ␯ϭ ϭ ϭ ϭ 2 , 31 1/4, p 3/8 , the fraction 2/9 for m 5, p 20, r 2 mpϪr ͑ ϭ ͒ ϭ ϭ ϭ ␯ ϭ ␯ ϭ or r 8 or for m 7, p 28, r 4( e 2 p 1/2), and the ␯ϭ ϭ ϭ ϭ ␯ ϭ ␯ ϭ where m is a positive odd number (mу3), p is an even fraction 2/11 for m 7, p 28, r 8( e 2 p 1/2). number (pу4), and r is a number of arbitrary parity (r Relation ͑31͒ also contains other experimentally ob- у2). Here the relative numbers of unpaired 2D electrons served fractional values of the filling factor which are not and bound ͑Cooper͒ pairs are given by contained in the composite fermion model. For example, the ␯ϭ ͑ ͒ ϭ ϭ ϭ Ϫ fraction 3/11 follows from 31 for m 3, p 4, r 10 Ne p 2r ␯ ϭ␯ ϭ ϭ ϭ ϭ ␯ ϭ ␯ ϭ ␯ ϭ ϭ ; ( e p 1/3) or for m 7, p 18, r 4( e 2 p 1/2), e N ϩ2N 4mϩpϪ4r ␯ϭ ϭ ϭ ϭ ͑␯ ϭ ␯ e p the fraction 4/11 for m 3, p 20, r 4 e 3/4, p ϭ ͒ ϭ ϭ ϭ ␯ ϭ ␯ ϭ ϭ N 2mϪr 1/8 or m 3, p 12, r 5( e 2 p 1/2), or for m 5, ␯ ϭ p ϭ ͑ ͒ ϭ ϭ ͑␯ ϭ ␯ ϭ ͒ ␯ϭ p ϩ ϩ Ϫ . 32 p 12, r 4 e 1/4, p 3/8 , and the fraction 3/13 for Ne 2N p 4m p 4r ϭ ϭ ϭ ␯ ϭ␯ ϭ m 7, p 20, r 6( e p 1/3). We note that as T de- Under the condition rϭ2m the Halperin relation ͑31͒ for creases in the Cooper phase, states with a higher concentra- any m and p ͑including the point pϭ2rϭ4m, at which an tion of bound electron pairs should be realized. indeterminacy of the type 0/0 appears͒ gives the simple frac- In addition, under the condition pϭ2rϭ4n (n tions ␯ϭ1/m (mϭ3,5,7,...). We note that the condition r ϭ1,2,3,...) relation ͑31͒ contains fractions with even de- ϭ2m, according to Eq. ͑32͒, corresponds to a state with zero nominators, ␯ϭ1/2, 1/4, 1/6, etc., which, as a rule, are not ͑␯ ϭ ␯ ϭ ͒ 3,4 concentration of Cooper pairs p 0, e 1 , i.e., a ‘‘nor- observed in experiments. If it is assumed that the condition 5 у (0) ␥(1)ϭ␥(2) ϭ mal’’ Laughlin liquid in the temperature region T Tc . 0 0 holds for the lowest Landau level (s 0), then in ␯ϭ ͑ ͒ (0) ϭ However, the fractions 1/m can also be contained in re- accordance with 29 we obtain TCDW 0 in the entire inter- ͑ ͒ ␯ Þ р␯ р (0)ϭ ␯ ϭ ͑ ͒ lation 31 for any finite values p 0, i.e., in the Cooper val 0 0 1 and Tc 0 at the point 0 1/2 see Fig. 9a . Ͻ (0) ␯ϭ phase T Tc . For example, the fraction 1/3 is realized, This means that in the FQHE regime in the neighborhood of 798 Low Temp. Phys. 27 (99–10), September–October 2001 E. A. Pashitski

␯ϵ␯ ϭ 0 1/2 the quantum states with a Cooper gap are ex- cluded and, consequently, those features of RH and Rxx which are determined by relation ͑31͒ should be absent, as should the features of Rxx and Ryy due to the quantum CDW. Nevertheless, a pseudogap at ␯ϭ1/2 has been observed in tunneling experiments.42 At the same time, because of the interelectron attraction due to the electron–phonon and electron–plasmon interac- tions with surface phonons and plasmons at sufficiently low temperatures, triplet Cooper pairing of composite fermions is, in principle, possible, and such pairing can be accompa- nied by the appearance of a quantum feature at ␯ϭ5/2.14 FIG. 10. Experimental phase diagram of a 2D system in the FQHE regime, obtained in Ref. 44 on the basis of the results of magnetooptic measure- ments on different samples: ϩ,᭹—Nϭ5.3ϫ1010 cmϪ2; ᭿—Nϭ5.2 5. DISCUSSION OF THE RESULTS AND COMPARISON WITH Ϫ Ϫ ϫ1010 cm 2; ᭡—Nϭ6.1ϫ1010 cm 2; ᮀ—results of Ref. 45; ᭝—results EXPERIMENT of Refs. 46 and 47. The mechanism proposed in this paper for the Cooper pairing of 2D electrons in a quantizing magnetic field, based ␯ on their interaction with 2D phonons and 2D plasmons lo- the study of nonlinear magnetotransport phenomena ( c calized near the interface of the crystals, enables one to de- ϭ0.22); this is most likely due to the accuracy of the corre- scribe the quantum features of the transport and Hall resis- sponding experimental methods. tances in the IQHE and FQHE regimes in a unified approach, In the ‘‘solid’’ phase, which in Refs. 43 and 44 was taking into account the power-law divergence of the Cooper correlated with the behavior of the a weakly damped line in and zero-sound diagrams at T→0. Summation of the diver- the spectrum of recombination radiation, quantum features ͑ ␯Ͻ␯ ͒ gent ladder diagrams of thermodynamic perturbation theory were observed in the FQHE regime including at c at in the zero-sound and Cooper channels in the local approxi- values of the filling factor ␯ϭ2/3, 2/5, 2/7, 2/9, 2/11, mation gives an expression for the vertex part with a pole at which are characteristic for the Cooper phase ͑see Sec. 4͒ a certain temperature which depends on the degree of filling and are contained in the Halperin relation ͑31͒. In this con- of the upper Landau level. This allows one to construct the nection one can assume that the weakly damped line in the phase diagram of a 2D system in a quantizing magnetic field; luminescence spectrum of the 2D system may be due not to in the case when the contribution of electron–hole processes Wigner crystallization but to Cooper pairing of 2D electrons, is dominant, this phase diagram contains both a Cooper and since the Cooper phase in the FQHE regime (␯Ͻ1) may be an excitonic phase ͑Fig. 8͒. The latter is localized near half- a precursor of the superconducting coherent phase in the integer values of the filling factor ␯ϭ(2nϪ1)/2, in qualita- IQHE regime (␯Ͼ1). tive agreement with the experimental data25 from observa- An additional argument in favor of this assumption is the tion of the anisotropic features of the transport resistance, qualitative similarity between the theoretical dependence of (0) which are due to the scattering of 2D electrons on quantum Tc (H) in Fig. 9a and the envelope of the phase boundaries CDWs.29–31 The absence of such features for nϭ1 and 2 in Fig. 10. As to the ‘‘fine structure’’ of the empirical phase may be due to enhancement of the contribution of electron– diagram, which is manifested in the existence of narrow dips electron processes on the lower Landau levels, when the on the field dependence of the critical temperature near the Cooper pairing of 2D electrons become the leading effect. fractional values ␯ϭ1/5,1/7, etc., a natural explanation for Here the superposition of bound electron pairs and unpaired this effect ͑as also for the case of the Wigner crystal͒43,44 is electrons in the Cooper phase in the ultraquantum limit gives that the energy of the ground state of the Laughlin liquid5 has rise to new features of the Hall and transport resistances at deep minima at ␯ϭ1/(2nϩ1), which lie below the energy fractional values of ␯, which are not described by the com- of the Cooper phase at the same values of ␯. posite fermion model. Thus it is not ruled out that what was observed in the It is interesting to compare the theoretical phase diagram magnetooptic experiments43,44 was not a ‘‘solid’’ Wigner for a system in the FQHE regime for the Cooper phase under phase but a ‘‘liquid’’ Cooper phase, arising because of the ␥(0)ϭ␥(0) ͑ ͒ the conditions 1 2 Fig. 9a with the experimental interelectron attraction due to the electron–phonon and phase diagram constructed in Refs. 43 and 44 on the basis of electron–plasmon interactions with surface phonons and ͑ ͒ ␯ϭ (0)Ϸ magnetooptic measurements Fig. 10 . This empirical phase plasmons. In the neighborhood of 1/2, where Tc 0 for ␥(0)Ϸ␥(0) diagram was interpreted in Refs. 43 and 44 as being the 1 2 , such attraction at low T may promote triplet result of Wigner crystallization of 2D electrons in a quantiz- Cooper pairing of composite fermions14 and the appearance ing magnetic field as T is lowered. However, as was noted in of a quantum feature at ␯ϭ5/2.11–13 Ref. 43, the measured temperature of the liquid–crystal tran- In closing, the author thanks Yu. A. Bychkov and I. B. sition turned out to be much lower than the classical melting Levinson, V. B. Timofeev, and I. V. Kukushkin for helpful temperature of the 2D Wigner crystal at the corresponding discussions. electron concentration. Furthermore, the tentative boundary This paper is dedicated to the centenary of the birth of of the existence region of the Wigner crystal indicated in the outstanding experimental physicist Lev Vasilievich ␯ ϭ Refs. 43 and 44 ( c 0.26) does not agree with the data from Shubnikov, who was among the earliest discoverers of quan- Low Temp. Phys. 27 (99–10), September–October 2001 E. A. Pashitski 799 tum effects in high magnetic fields, of which the quantum 24 P. Phillips, S. Knysh, and D. Davidovich, Nature ͑London͒ 395,253 Hall effect in 2D systems is an example. ͑1998͒. 25 M. R. Lilly, R. B. Cooper, J. P. Eisenstein, L. N. Pfeiffer, and K. W. West, Phys. Rev. Lett. 82, 394 ͑1999͒. * E-mail: [email protected] 26 H. Fukuyama, P. M. Platzman, and P. W. Anderson, Phys. Rev. B 19, 5211 ͑1979͒. 27 ´ Yu. A. Bychkov, S. V. IordanskiŽ, and G. M. Eliashberg, JETP Lett. 33, ͑ ͒ ͑ ͒ 1 K. Von Klitzing, C. Dorda, and M. Pepper, Phys. Rev. Lett. 45, 494 143 1981 ; Poverkhnost’ 10,33 1982 . 28 ´ ͑ ͒ ´ ͑1980͒. Yu. A. Bychkov and E. I. Rashba, JETP Lett. 38, 229 1983 ; Zh. Eksp. ͑ ͓͒ ͑ ͔͒ 2 D. C. Tsui, H. L. Stormer, and A. C. Gossard, Phys. Rev. Lett. 48, 1559 Teor. Fiz. 85, 1826 1983 Sov. Phys. JETP 58, 1062 1983 . 29 ͑1982͒. A. A. Kulakov, M. M. Fogler, and B. I. Shklovskii, Phys. Rev. Lett. 76, ͑ ͒ 3 Kh. Shtermer, Usp. Fiz. Nauk 170,304,͑2000͒. 499 1996 . 30 4 D. Tsui, Usp. Fiz. Nauk 170, 320, ͑2000͒. M. M. Fogler, A. A. Kulakov, and B. I. Shklovskii, Phys. Rev. B 54, 1853 5 R. B. Laughlin, Phys. Rev. Lett. 50,1395͑1983͒; Science 242, 525 ͑1996͒. 31 ͑1988͒; Usp. Fiz. Nauk 170,294͑2000͒. R. Moess and J. T. Chalker, Phys. Rev. B 54, 5006 ͑1996͒. 32 6 F. D. M. Haldane, Phys. Rev. Lett. 51, 605 ͑1983͒. B. I. Halperin, Helvetika Physica Acta 56,75͑1983͒. 7 B. I. Halperin, Phys. Rev. Lett. 52, 1583 ͑1984͒. 33 Yu. A. Bychkov, Fiz. Tverd. Tela ͑Leningrad͒ 31͑7͒,56͑1989͓͒Sov. Phys. 8 F. D. M. Haldane and E. H. Rezayi, Phys. Rev. Lett. 54, 237 ͑1985͒. Solid State 31, 1130 ͑1989͔͒. 9 ͑ ͒ 34 J. K. Jain, Phys. Rev. Lett. 63, 199 1989 . V. F. Gantmakher and I. V. Levinson, Rasseyanie nositeleŽ toka v metal- 10 B. I. Halperin, P. A. Lee, and N. Read, Phys. Rev. B 47, 7312 ͑1993͒. lakh i poluprovodnikakh, Nauka, Moscow ͑1984͒. 11 35 ´ ͑ ͒ E. Mendez, L. I. Chang, C. A. Chang, L. F. Alexander, and L. Esaki, Surf. L. G. Il’chenko and E. A. PashitskiŽ, Fiz. Tverd. Tela Leningrad 22,3395 Sci. 142,215͑1984͒. ͑1980͓͒Sov. Phys. Solid State 22,1987͑1980͔͒. 12 R. Willett, J. P. Eisenstein, H. L. Stormer, D. C. Tsui, A. C. Gossard, and 36 A. V. Andreev and Yu. A. Bychkov, Phys. Rev. B 45, 1443 ͑1992͒. ͑ ͒ 37 ´ J. H. English, Phys. Rev. Lett. 59,1776 1987 . Yu. A. Bychkov, L. P. Gor’kov, and I. E. DzyaloshinskiŽ, Zh. Eksp. Teor. 13 W. Pan, J. S. Xia, V. Shvarts, D. E. Adams, H. L. Stormer, D. C. Tsui, Fiz. 50, 738 ͑1966͓͒Sov. Phys. JETP 23,489͑1966͔͒. 38 ´ ͑ ͒ L. N. Pfeiffer, K. W. Baldein, and K. W. West, Phys. Rev. Lett. 83, 3530 I. E. DzyaloshinskiŽ and A. I. Larkin, Zh. Eksp. Teor. Fiz. 61, 791 1971 ͑1999͒. ͓Sov. Phys. JETP 34, 422 ͑1972͔͒. 14 ͑ ͒ ͑ ͒ 39 ´ ͑ ͒ V. W. Scarola, K. Park, and J. K. Jain, Nature London 406,863 2000 . L. P. Gor’kov and I. E. DzyaloshinskiŽ, Zh. Eksp. Teor. Fiz. 67,397 1974 15 ´ ͑ ͒ ͓ ͑ ͔͒ A. L. Kasatkin and E. A. PashitskiŽ, Ukr. Fiz. Zh. Russ. Ed. 22,467 Sov. Phys. JETP 40, 198 1975 . ͑1977͒. 40 A. T. Zheleznyak, V. M. Yakovenko, and I. E. Dzyaloshinskii, Phys. Rev. 16 F. D. M. Haldeine and E. Rezayi, Phys. Rev. Lett. 60, 956 ͑1988͒. B 55, 3200 ͑1997͒. 17 T. Maniv, A. I. Room, I. D. Vagner, and P. Wyder et al., Phys. Rev. B 46, 41 J. R. Schrieffer, Theory of Superconductivity ͓Benjamin, New York 8360 ͑1992͒. ͑1964͒; Nauka, Moscow ͑1970͔͒. 18 E´ . A. Pashitskii, Fiz. Nizk. Temp. 25, 920 ͑1999͓͒Low Temp. Phys. 25, 42 K. M. Brow, N. Turner, J. T. Nicholls, E. H. Linfield, M. Pepper, D. A. 690 ͑1999͔͒. Ritchie, and G. A. C. Jones, Phys. Rev. B 50, 15465 ͑1994͒. 19 Z. V. Popovic, M. Cardona, E. Richter, D. Strauch, L. Tapfer, and 43 I. V. Kukushkin and V. B. Timofeev, Usp. Fiz. Nauk 163,1͑1993͒. K. Ploog, Phys. Rev. B 41,5904͑1990͒. 44 I. V. Kukushkin and V. B. Timofeev, Adv. Phys. 45,147͑1996͒. 20 C. Trallero-Ginev, F. Garcia–Moliner, V. R. Velasco, and M. Cardona, 45 V. I. Goldman, M. Santos, M. Shayagan, and J. E. Cunnigham, Phys. Rev. Phys. Rev. B 45, 11944 ͑1992͒. Lett. 65, 2198 ͑1990͒. 21 D. Simonyan, S. V. Kravchenko, M. P. Sarachik, and V. M. Pudalov, Phys. 46 M. A. Paalanen, R. L. Willett, P. Littlewood, R. P. Ruel, K. W. West, L. N. Rev. Lett. 79, 2304 ͑1997͒. Pfeiffer, and D. J. Bishop, Phys. Rev. B 45, 11342 ͑1992͒. 22 ´ ´ ͑ ͓͒ 47 E. A. PashitskiŽ, Zh. Eksp. Teor. Fiz. 56, 662 1969 Sov. Phys. JETP 29, M. A. Paalanen, R. L. Willett, P. R. Ruel, P. Littlewood, K. W. West, and 362 ͑1969͔͒. L. N. Pfeiffer, Phys. Rev. B 45, 11784 ͑1992͒. 23 ´ A. M. Gabovich, D. P. Moiseev, E. A. PashitskiŽ, and S. K. Uvarova, Fiz. Nizk. Temp. 1,984͑1975͓͒Sov. J. Low Temp. Phys. 1,473͑1975͔͒. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Bound spin-excitons in two-dimensional electron gas V. Fleurov*

School of Physics and Astronomy, Beverly and Raymond Sackler Faculty of Exact Sciences, Tel Aviv University, Tel Aviv 69978, Israel K. Kikoin

Physics Faculty, Ben-Gurion University of the Negev Beer-Sheva 84105, Israel ͑Submitted April 23, 2001͒ Fiz. Nizk. Temp. 27, 1083–1091 ͑September–October 2001͒ A theory of the spin exciton capture by a magnetic impurity in a 2D electron gas is developed. We consider a resonance model for the electron scattering by a transition metal impurity and calculate the binding potential for spin excitons. This potential is spin selective and is capable of binding a spin exciton with zero angular momentum. In order to trap an exciton with a nonzero angular momentum m, the potential must exceed a certain threshold value that depends on m.©2001 American Institute of Physics. ͓DOI: 10.1063/1.1414568͔

1. INTRODUCTION ions. The intra-atomic exchange interaction leads to Hund’s rule, which governs the occupation of the deep levels. As a The Shubnikov–de Haas effect is a powerful tool for result, the transition atoms in a semiconductor are magnetic. ͑ ͒ studies of a two-dimensional electron gas 2DEG in a strong We will show below that the interplay between the magnetic magnetic field. In particular, the oscillatory behavior of the impurity scattering and attractive electron-hole interaction in electronic g factor in a 2DEG has been investigated with the excited 2DEG results in a bound spin exciton. The spectrum ͑ ͒ help of this effect see Ref. 1 and references therein .Two of these bound states is the subject of the present study. important parameters which predetermine the properties of the 2DEG in a strong magnetic field B are the cyclotron ␻ ϭ frequency B eB/mc and the effective Coulomb energy 2. MODEL AND APPROXIMATIONS ϭ 2 ␬ ϭ ប 1/2 Ec e / lB (lB ( c/eB) ) is the magnetic length. In the Ӷប␻ We start with the model of a 2DEG doped by magnetic limit of ultrahigh magnetic fields, when Ec B and only the lowest Landau sublevel is filled, the low-energy branches impurities which is discussed in detail in Ref. 10. This model of the excitation spectrum are represented by well-separated is described by the Hamiltonian ϭ ϩ ͑ ͒ bands of spin waves, magnetoplasmons, etc. These excita- H Hb Hi 1 tions have been studied in detail on recent decades ͑see, e.g., where Refs. 2–7͒. Various forms of spin excitations can be ob- served experimentally using the inelastic light scattering ϭ ϩ ͑ ͒ Hb ͚ En␴anm,␴anm,␴ 2 method as described in Ref. 8, where features are observed nm␴ that are attributed to collective excitation in a 2DEG. The presence of magnetic impurities may account for additional is the Hamiltonian of a 2DEG strongly quantized by a mag- features due to localized spin excitations. netic field. It is convenient to use the symmetric cylindrical ϭ Ϫ The spin waves formed by the electrons in the down- gauge for the vector potential A (( B/2)y,B/2,0), so that ͑ ͒ spin and holes in up-spin nϭ0 Landau subbands make up the index m in 2 describes different orbital states in a given the lowest branch of magnetic excitations. These states are Landau level ⌬ ϭ ␮ 2 separated by the Zeeman gap B g BB from the ground 1 ប 2nϩ1 ϭប␻ ͩ ϩ ͪ ϵ ͑ ͒ state and form a band with a width of ϳE . The exciton En B n . 3 c 2 2m l2 dispersion law is quadratic at small wave numbers and satu- B rates in the short-wave length limit, where the excitons, in Then, assuming a small impurity concentration, the im- 2 ͑ ͒ fact, transform into free electron-hole pairs. purity related part Hi of the Hamiltonian 1 can be written This paper studies the interaction between spin waves in the general form and a magnetic impurity in a 2DEG. To be more specific, we ϭ ͗ ⌫͉ ͉ ⌫͘ϩ ͑͗ ␥͉ ͉ ͘ ϩ consider doped heterojunctions GaAs/GaAlAs and related Hi ͚ i E⌫ i ͚ ͚ i V nm di␥␴anm␴ materials, so we refer below to a III–V semiconductor as a ⌫ ␥ nm␴ host material, in which some cation atoms are substituted for 9 ϩ ͒ϩ ͗ ͉⌬ ͉ Ј Ј͘ ϩ ͑ ͒ magnetic transition metal impurities. It is known that tran- h.c. ͚ nm V n m anm␴anЈmЈ␴ . 4 ␴ sition metal atoms create deep levels in the forbidden energy nm gap of the host semiconductor, and the main mechanism of Here the state of an isolated impurity ion is characterized by the electron–impurity scattering is the resonance scattering a configuration dn of its unfilled d shell in a crystal field by the d levels of the unfilled 3d shell of transition metal preserving the point symmetry of the bulk semiconductor

1063-777X/2001/27(9–10)/7/$20.00800 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 V. Fleurov and K. Kikoin 801

͑we assume that the potential responsible for the confinement in the z direction does not perturb the crystalline environ- ment of the impurity site͒. Then the electrons in the d shell ␥ϭ are characterized by the representations t2 ,e of the tetra- hedral point group, and the many-electron states ͉⌫͘ of the 3d shell may be represented as

n r1 r2 r3 r4 d⌫ϭ͑ne e t t ͚͒ ϭ ↑ ↓ 2↑ 2↓ ri n i ͑↓ and ↑ are two projections of the electron spin͒. The scat- 9 tering part of Hi consists of two components: the second term in Eq. ͑4͒ describes the resonance part of the impurity scattering; the third term represents the short-range substitu- ⌬ ϭ Ϫ Ϫ Ϫ tion potential V Vi(r R0) Vhost(r R0), where R0 is the position of the substitutional impurity in the host lattice. FIG. 1. A schematic representation of the interacting Landau and impurity The resonance scattering arises together with the usual ‘‘po- levels. Eie␴ are the impurity levels for spin-up and spin-down electrons. ␧ ␧ They result from the prime impurity levels ie␴ shifted to new positions due tential’’ scattering due to the fact that the energy level i␥ ϭ n nϪ1 to their interaction with the Landau levels with m 0. The Landau levels ϭE⌫(d )ϪE⌫ ,(d ) enters the fundamental energy gap of ϭ with m 0 are also shifted to their new positions Ebe␴ . The inset illustrates the host semiconductor or appears in the lowest conduction the case when an impurity prime spin-down state is nearly degenerate with band of the topmost valence band. Here the configuration the lowest Landau spin-down state. nϪ1 ␥ϭ d is missing one electron in a state e or t2 in com- parison with the state dn. The impurity problem with the two scattering mecha- This observation allows us to divide the states in the nisms can be solved in the general case.11,12 In our special Landau band ͑2͒ into two groups, case its solutions are essentially different for the e and t2 channels. As is shown in Ref. 10, the t component of the ϭ ϩ Ј ͑ ͒ 2 Hb Hb0 Hb . 8 impurity potential results in deep levels in the forbidden en- ergy gap or resonances in the conduction band, but only Here Hb0 includes only the states with the zero angular mo- ment mϭ0, whereas HЈ includes all the remaining states. weakly perturbs the Landau levels En . On the other hand, b e-scattering results in the appearance of bound Landau states Only the states with mϭ0 are involved in the formation of with mϭ0 between Landau levels with nonzero orbital the bound Landau states given by the solutions of Eq. ͑6͒. quantum number and in a splitting of the lowest bound state Similar reduction can be made in the manifold ͕i␥␴͖ due to from the Landau grid. The reason for this difference is in the the ‘‘selection rule’’ ͑5͒: when considering the renormaliza- short-range nature of both components of the scattering po- tion of the states ͉␤͘ due to the resonance scattering, we tential. The eventual reason for this difference is the orbital retain only the state ͉e1͘ in the corresponding sector of the dependence of the matrix element secular matrix. As a result, one has the following equation ␳ ͉m͉ ͗ ␴͉⌬ ͉ ␴͗͘ ␴͉⌬ ͉ ␴͘ ͑0͒ bn U e1 e1 U bn ͗ ͉⌬ ͉␥␮͘ϳͩ ͪ Ӷ ͑ ͒ E ␴ϪE ϭ nm U 1 5 bn bn␴ Ϫ lB Ebn␴ Eie␴ 9 ϩ␦ ͒ ͑ ͒ for mÞ0. Here ␳ is the radial variable in the cylindrical M͑Ebn␴ 9 ␳ ␸ coordinates ( , ,z). When estimating this matrix element (0) one should choose ␳ of the order of the atomic radius. As a for the energy shift of a given level Ebn␴ which according to ͉ ϰ͉ 2Ϫ 2 ϳ ϭ Ref. 13 appears in the gap between the bare Landau levels result, only the orbitals e1͘ r 3z ͘ Y 20 with m 0 may be strongly hybridized with the Landau states. EnϪ1 and En . Here we have picked up the direct mutual According to Refs. 10 and 12, the localized eigenstates repulsion of the nth Landau level and the d level in the first ͑ ͒ of the Hamiltonian ͑1͒, ͑4͒ are given by the following equa- term of the right-hand side of Eq. 9 . The influence of the tion other bound Landau states is given by the second term, ͗␤Ј͉⌬ ͉ ␴͗͘ ␴͉⌬ ͉␤Ј͑͘ Ϫ ͑0͒͒ E ␥␴Ϫ␧ ␥␴ϪM ␥͑E ␥␴͒ϭ0, ͑6͒ U e1 e1 U E E␤ i i i ␦M͑E͒ϭ ͚ . ͑ Ϫ ͒͑ Ϫ ͑0͒͒ where ␤ЈÞ␤ E Eie␴ E E␤Ј ͗␥␮͉⌬U͉␤͗͘␤͉⌬U͉␥␮͘ The role of these states is to keep the renormalized levels in ͑ ͒ϭ ͑ ͒ M ␥ Ei␥␴ ͚ , 7 the same energy interval E Ϫ ␴ϽE ␴ϽE ␴ . ␤ E ␥␴ϪE␤ n 1, bn n i An important feature of the resonant channel of impurity where ͉␤͘ϵ͉bn␴͘ stands for the eigenfunctions of the mag- scattering is its spin selectivity, which stems from the spin netically quantized electrons captured by the local potential structure of the transition metal d shell. It is known9 that in ⌬U. The latter problem was solved in Ref. 13 for an attrac- their main features the transition metal impurities follow the tive potential ⌬UϽ0. It was shown that the short-range po- ‘‘Aufbau principle’’ of the quantum mechanics of isolated tential also perturbs only the zero orbital states mϭ0. The atoms. This means that the d shell is usually filled in accor- general structure of the electron spectrum of a doped 2DEG dance with Hund’s rule or, in other words, the exchange is sketched in Fig. 1. interaction makes the level Eie↑ lie always below the level 802 Low Temp. Phys. 27 (99–10), September–October 2001 V. Fleurov and K. Kikoin

ϩ E ↓ . Then the splitting ⌬ ϭE ↓ϪE ↑ generates the spin ϭ͉͑ ͉2 Ϫ ͒ ie es ie ie Hi A00 Eb0↑ E0↑ ai0ai0 splitting ⌬(n) of the bound Landau states via Eq. ͑9͒. This bs ϩ͉͑ ͉2 Ϫ ͒ ϩ ϩ ͑ ͒ splitting can be estimated as B00 Eb0↓ E0↓ bi0bi0 Hi,nd . 13 2 ͑ ͒ ⌬ ͉V ͉ ␦ Here Eb0␴ are the solutions of Eq. 9 for the lowest bound ͑n͒ es 2 enЈ n ϩ ϩ ⌬ Ϸ ͑ ͒ ͑ ͒ ͯV ͯ ϩ ͚ ͑ ͒ . ͑10͒ ͩ er ͑ ͒ ͪ bs n n 0 0 Landau state, and the operators ai0 and bi0 create the corre- ⌬↑ ⌬↓ Þ E ϪE nЈ n bn bnЈ sponding eigenfunctions in the symmetric cylindrical gauge. ⌬(n)ϭ Ϫ ϭ͗ ͉⌬ ͉ ͘ ␦ ϭ Ϫ (0) The coefficients A ,B are defined in Eq. ͑A2͒. The off- Here ␴ Ebn Eie␴ , Ven e1 U bn , n Ebn Ebn 00 00 ͑the spin dependence of the two last quantities is neglected͒. diagonal term Hi,nd contains the contribution from the higher It is important for the further classification of the spin exci- Landau levels with nϾ0, whose value is of the order of ⌬(n) ⌬ ϳ͉A ͉2/E Ӷ1 and will be neglected in the further calcula- tons that the sign of bs , unlike the sign of es , can be both 0n n (n) negative or positive, since the energy differences ⌬↑ and tions. (n) In principle, the scattering potential also contains terms ⌬↓ can have either the same or the opposite signs depend- ing on the type of the transition metal ion and the host matrix corresponding to the spin-flip processes, ͑see the discussion in Ref. 10͒. According to Eq. ͑10͒ the ϭ ͑ ϩ ϩ ϩ ϩ ϩ ͒ ͑ ͒ HЌ J die↑die↓bi0ai0 die↓die↑ai0bi0 . 14 resonance impurity scattering results in a ‘‘transfer’’ of the exchange splitting ⌬ of the impurity d shell to the spectrum The spin-flip terms are inessential in comparison with the es ͑ ͒ of the Landau electrons. This transfer is illustrated by the leading spin-diagonal terms 13 because each spin flip costs ⌬ ϭ Ϫ ϳ͉ ͉2 ⌬ level renormalization in Fig. 1. the Hund energy es Eie↓ Eie↑ , so that J Ve0 / ex . Now we know the general structure of the one-particle The spin-flip processes, in principle, result in a multiple cre- spectrum of a magnetically doped 2DEG, which should serve ation of spin excitons, but the contribution of these processes ϳ 2 as a background when magnetic excitons and magnetoplas- to the spin-wave spectrum is at least J , and we also ne- mons are formed. This spectrum consists of equidistant Lan- glect them in the further calculations. dau levels with mÞ0 which are unperturbed by the impurity Then, we turn from the cylindrical gauge to the Landau scattering. The states with mϭ0 form their own grid: a pair gauge, (0) of spin-split states Ebn␴ appears in each energy gap En ϩ ϩ ϩ ϩ (0) ϭ ϭ ͑ ͒ ϪE Ϫ , and the lowest pair of bound states E ␴ arises in ai0 ͚ A pa p , bi0 ͚ B pb p , 15 n 1 bn p p the fundamental energy gap of the 2D semiconductor below the Landau level E0 . where ϭ ↑͉ ↑ ϭ ↓͉ ↓ ͑ ͒ A p ͗i0 p ͘, B p ͗i0 p ͘. 16 3. ENERGY SPECTRUM OF LOCALIZED SPIN EXCITONS As a result the impurity Hamiltonian in the Landau According to the general classification of the multipar- gauge has the form ticle excitations from a filled Landau level in a 2DEG,4 the ϭ ͓ ͑ Ј͒ ϩ ϩ ͑ Ј͒ ϩ ͔ ͑ ͒ lowest branch of the electron–hole excitations is that of the Hi ͚ U↑ p,p a p a pЈ U↓ p,p b p b pЈ 17 spin waves. These spin excitons arise as bound electron–hole ppЈ pairs with parallel spins as a result of a spin-flip excitation where from the lower filled sublevel E ↑ of the lowest Landau level 0 ͒ϭ ͒ϭ ͑ ͒ U↑͑p,pЈ K↑J p↑J* ↑ , U↓͑p,pЈ K↓J p↓P* ↓ . 18 to its higher empty sublevel E0↓ . To calculate the spectrum pЈ pЈ of the localized spin excitons, one should keep only the Here J ␴ϭ͗⌿ ␴͉⌿ ␴͘, and ⌿ ␴ are the wave functions ϭ ͑ ͒ p b p,0 p,0 states with n 0 in the bare Hamiltonian Hb 2 , take into of the lowest Landau level, account the Zeeman splitting of the Landau states explicitly, 1 ͑xϩp͒2 and then add the electron-hole Coulomb interaction Hint , ␺ ͑ ͒ϭ ͑ ͒ ͩ Ϫ ͪ ͑ ͒ p,0␴ x,y exp ipy exp . 19 ͑2␲3/2͒1/2 2 0 ϭ ͑ ϩ ϩ ϩ ͒ϩ ͑ ͒ Hex ͚ E0↑a p a p E0↓b p b p Hint . 11 p Here x,y are Cartesian projections of the dimensionless vec- tor r/l . The coefficients K␴ are calculated in the Appendix. Operators aϩ(a) and bϩ(b) describe creation ͑annihilation͒ B To derive the impurity Hamiltonian in the Landau gauge, of electrons with the up or down spins ␴, index nϭ0 being we use the identity omitted. Here the electron wave functions are written in the asymmetric Landau gauge AϭB(y,0,0). Then the total dq pϩpЈ ϩ ϭ ͵ x ͫϪ ͬ␳ ͑ ЈϪ ͒ a a exp iq ↑ q ,p p , Hamiltonian of a doped 2DEG takes the form p pЈ 2␲ x 2 x ϭ 0 ϩ ͑ ͒ H Hex Hi . 12 dq pϩpЈ ϩ ϭ ͵ x ͫϪ ͬ␳ ͑ ЈϪ ͒ ͑ ͒ b p b pЈ ␲ exp iqx ↓ qx ,p p , 20 The impurity-related term Hi was analyzed in the previous 2 2 Section using the symmetric gauge. Now it should be re- where the electron density operators are derived in the Landau gauge. q Using the solution of the impurity problem found in the ␳ ͑ ͒ϭ ϩ ͩ ͩ ϩ y ͪͪ ↑ q ͚ a p a pϩq exp iqx p previous Section and taking into account that the Landau p y 2 ϭ levels with m 0 are included in Hex , contrary to the ‘‘ex- ͑ ͒ ϩ qy traction’’ principle formulated in Eq. 8 , we can write Hi in ␳ ͑ ͒ϭ ͩ ͩ ϩ ͪͪ ͑ ͒ ↓ q ͚ b p b pϩq exp iqx p 21 the form ͑see Appendix͒ p y 2 Low Temp. Phys. 27 (99–10), September–October 2001 V. Fleurov and K. Kikoin 803

with qϭ(q ,q ). Then inserting Eq. ͑20͒ into Eq. ͑17͒,we x y ͓␧ ͑ ͒Ϫ␧͔ ͑ ͒ϩ ˜ ͑ Ϫ Ј͒ get, after straightforward calculations, ex k f k 2i͚ Uϩ k k kЈ

2 1 ϭ ͑ ͒ q /4␳ ͑ ͒ ͑ ͒ ϫ ϫ ͒Ϫ ˜ Ϫ ͒ Hi ͚ U␴ q e ␴ q , 22 sin ͓kЈ k͔z f ͑kЈ 2͚ UϪ͑k kЈ q,␴ 2 kЈ where the matrix elements of the impurity potential are 1 ϫcos ͓kЈϫk͔ f ͑kЈ͒ϭ0, ͑28͒ 2 z ͑ ͒ϭ ͵ 2 2 ␺ for the envelope function f (k). Here U␴ q K␴ d r1d r2 b␴ 1 2 2 Ϫq2/4 Ϫq2/4 r r ͉r Ϫr ͉2 U˜ Ϯ͑q͒ϭ ͑U↑͑q͒ϮU↓͑q͒͒e ϵUϮ͑q͒e . ϫͩ 1 ͪ ␺ ͩ 2 ͪ ͩ Ϫ 1 2 ͪ 2 ␴ exp 2 b 2 4 As is discussed in the Appendix, the localization radius ␳ 1 b of the impurity wave function is substantially smaller than ϫexpͭ ͓͑x Ϫx ͒q ϩ͑y Ϫy ͒q Ϫi͑y 2 1 2 y 1 2 x 1 the magnetic length. Then the q dependence of the matrix elements ͑23͒ is insignificant, and they can be estimated as ϩ ͒ ϩ ͑ ϩ ͒ Ϫ ͑ ϩ ͒͑ Ϫ ͔͒ͮ ϱ 2 y 2 qy i x1 x2 qx i x1 x2 y 1 y 2 . 2 U␴ϷK␴͑2␲ ͒ͩ ͵ d␰␺ ␴͑␰͒ͪ ϵK␴I . ͑29͒ b b 0 ͑23͒ Taking into account the cylindrical symmetry of the problem, ͑ ͒ Here the coefficients K␴ determined in the Appendix depend we seek the solutions of Eq. 28 in the form on the specific form of the scattering impurity potential act- ͒ϭ ͒ im␸ ͑ ͒ f ͑k f m͑k e , 30 ing on the electrons in the Landau subband 0␴ ͓see Eqs. ͑16͒, ͑A3͒, ͑A6͒, ͑A8͔͒. We discuss here the limit of a strong mag- where the integer quantity m is the quantum number of the netic field, when the cyclotron energy ប␻ ϭប2/2ml is bound exciton. B B ͑ ͒ ͑ ͒ large compared to the Coulomb energy e2/␬l . It is essential Now we substitute the functions 30 in Eq. 28 and B Ј that the impurity potential is spin selective, i.e., its compo- carry out the integration over the directions of the vector k . nents acting on the electrons in the two Landau subbands can Then the term proportional to Uϩ contains the integrals differ significantly in magnitude ͑see below͒. 2␲ d␸ kkЈ kkЈ We consider the case of the filling factor vϭ1, when the ͵ expͩ cos ␸ͪ sinͩ sin ␸ͪ sin m␸ 0 2␲ 2 2 spin-up Landau band is totally full and the spin-down Lan- dau band is completely empty. Then the eigenfunctions of ͑kkЈ͉͒m͉ ͑ ͒ ϭ ͉ ͉ϩ sign͑m͒ ͑31͒ the Hamiltonian 11 , 2 m 1͉m͉ k in which the sign function is defined as ⌿͑0͒ ϭ ϩ ͩ ͩ ϩ y ͉ͪͪ ͘ ͑ ͒ ex,k ͚ b p a pϩk exp ikx p 0 , 24 ␳ y 2 1, mϾ0 sign͑m͒ϭͭ 0, mϭ0, ͑32͒ correspond to the free spin-excitons with the energy spec- ␧ Ϫ1, mϽ0. trum ex(k), The term proportional to UϪ contains the integrals H͑0͒⌿͑0͒͑k͒ϭ␧ ͑k͒⌿͑0͒ . ͑25͒ ex ex ex,k 2␲ d␸ kkЈ kkЈ ͵ expͩ cos ␸ͪ sinͩ sin ␸ͪ cos m␸ ϭ 0 2␲ 2 2 Here k kx ,ky is the wave vector of a spin exciton. The exciton dispersion law is ͑kkЈ͉͒m͉ ϭ ͑ ͒ ͉m͉ϩ1͉ ͉ 33 e2 ␲ 1/2 2 m ␧ ͑ ͒ϭ⌬ ϩͩ ͪͩ ͪ ͓ Ϫ Ϫk2/4 ͑ 2 ͔͒ ex k B ␬ 1 e I0 k /4 lB 2 for all values of the quantum number m. The equations defining the radial parts f (k) of the en- ϵ⌬ ϩ⍀͑ 2͒ ͑ ͒ m B k 26 velope functions ͑30͒ are ͑ ͒ ⌬ ϭ͉ ␮ ͉ 1 see Refs. 2–4 and 14 . Here B g BB is the Zeeman ͑ ͒ϭ ͉m͉ Ϫk2/4 ͑ ͒ f m k Wnk e ␧Ϫ␧ ͒ Fm , 34 energy; I0 is a modified Bessel function. ex͑k The wave function of a bound exciton is looked for in where the form 1 ϭ ͕ ͓ ϩ ͑ ͔͒Ϫ ͓ Ϫ ͑ ͔͖͒ Wm ͉m͉ϩ1͉ ͉ U↓ 1 sign m U↑ 1 sign m ⌿ ϭ ͑ ͒⌿͑0͒ ͑ ͒ 2 m ! ex ͚ f k ex,k. 27 k and

The function ͑27͒ must be an eigenfunction of the Hamil- ϭ Ϫk2/4 ͉m͉ ͒ Fm ͚ e k f m͑k . tonian ͑12͒. The standard procedure leads to the equation k 804 Low Temp. Phys. 27 (99–10), September–October 2001 V. Fleurov and K. Kikoin

͑ ͒ ϭ ␧ ␧ϭ⌬ FIG. 2. A graphical solution of Eq. 35 for m 0. Due to the logarithmic divergence of the function M 0( )at B , a spin exciton can be always bound ␧ ͑ ͒ ͑ ͒ ϭϪ ␧ with an energy 0 below the bottom of the spin-exciton band a . A graphical solution of Eq. 35 for m 1. The function M Ϫ1( ) is now nondivergent. ␧ Hence, a spin exciton can be bound with an energy Ϫ1 below the bottom of the spin-exciton band if the parameter WϪ1 is larger than the threshold value ¯ ϭ ⌬ ͑ ͒ WϪ1 1/M Ϫ1( B) b .

␧ The energy m of the bound exciton with the quantum ⌬ ϭϪ ϭϪ ͑␤ Ϫ␤ ͒ϭ 2 es Ͼ ͑ ͒ number m can be found as a solution of the equation W0 UϪ VebIb ↑ ↓ VebIb 0. 39 ⌬↑⌬↓ ϭ ␧ ͒ ͑ ͒ 1 WmM m͑ m , 35 ͑ ͒ ͑ ͒ where The scattering potential 39 is repulsive and Eq. 35 for mϭ0 has no bound solutions below the exciton band. 2 k2͉m͉eϪk /2 Such a solution appears when the resonance scattering ͑␧͒ϭ ͑ ͒ M m ͚ ␧Ϫ␧ ͒ . 36 for the spin-down electrons is strong ͑see inset in Fig. 1͒. k ex͑k This is a realistic condition for transition metal impurities.8 2 We introduce now the new variable ␻ϭk and convert In that case we use Eq. ͑A7͒ for the potential acting on the the summation in Eq. ͑36͒ to integration. Then the fact that at spin-up electrons, and Eq. ͑A8͒ for the potential acting on small ␻ the dispersion law ⍀(␻)ϰ␻ allows us to find the the spin-down electrons. The coupling constant is ␧ ␧→⌬ behavior of the quantity M m( )at B . Ϫ␻ D ␻me /2 1 ͑␧͒ϭ␲ ͵ ␻ W ϭϪUϪϷϪ ͑V Ϫw ͒I . ͑40͒ M m d ␧Ϫ⌬ Ϫ⍀ ␻͒ 0 eb 0 b 0 B ͑ 2

Ϫ2␲m* ln m*͉⌬ Ϫ␧͉, mϭ0 ␧ Ϸ ex ex B Due to the logarithmic divergence of the function M 0( ) ͭ ͑⌬ ͒Ϫ Ј ͑⌬ ͉͒␧Ϫ⌬ ͉ ␧→⌬ Ϫ Þ . M m B M m B B , B 0, m 0 near the bottom of the excitonic band ͑see Eq. ͑37͒͒, a dis- ͑ ͒ ͑37͒ crete solution of Eq. 35 appears, provided the resonance component of impurity scattering is stronger than the poten- ϭ ␬ប2 2 ␲ប 1/2 Here mex* 2 /e (2eB/ c) is the effective mass of tial component, V Ͼw , i.e., W Ͻ0 ͑see Fig. 2͒. 2 ⌬ Ͻ Ј ⌬ eb 0 0 the free exciton at small momenta, M m( B) 0, M ( B) The selectivity with respect to the orbital quantum num- ϭ(dM(␧)/d␧)͉␧Ϫ⌬ →Ϫ Ͼ0. B 0 ber m of the bound spin exciton is intrinsically connected Now using the normalization condition with the spin selectivity of the impurity potential. The orbital momentum is the sum of the electron and hole momenta, ͉ ͉͒2ϭ ϭ ϩ ͚ f m͑k 1, m me mh . As is shown in Refs. 10 and 13, only the elec- k ϭ trons or holes with me,h 0 can be captured by the short- one finds that range impurity potential, which is U↓ for electrons and ϪU↑ for holes. Therefore, in a bound exciton with mÞ0 one of F ϭ͉W ͉͓MЈ͑␧ ͔͒Ϫ1/2. ͑38͒ m m m the carriers ͑electron or hole͒ must have zero momentum. This equation closes the procedure. Now the spectrum of Then the momentum of the whole exciton is, in fact, the spin excitations in 2DEG pinned by the resonance impurity momentum of the second carrier. This second carrier is with its own localized spin is determined by Eqs. ͑35͒–͑37͒. bound by the combined action of the Coulomb attraction of The coupling constants are determined in Eqs. ͑29͒, ͑A7͒, the first carrier and the diamagnetic contribution of the Lo- ͑A8͒, and the wave function of the bound spin exciton is renz force. The sign of the orbital momentum and the charge given by Eqs. ͑27͒, ͑34͒, and ͑38͒. of the carrier predetermines the sign, attractive or repulsive, Starting the analysis of the bound exciton solutions with of the Lorenz force contribution to the total potential. As a ϭ Ͼ the m 0 case, we see that in the weak-scattering limit the result, only the electrons with me 0 and the holes with ͑ ͒ Ͻ magnitudes of both coupling constants 29 are determined mh 0 can be captured in the limit of a strong magnetic field. by the coefficient ͑A7͒. Then the potential in Eq. ͑35͒ is Hence, the sign of the exciton orbital momentum provides given by the following equation information on which carrier is bound by the short-range Low Temp. Phys. 27 (99–10), September–October 2001 V. Fleurov and K. Kikoin 805

potential. It is the hole, bound by the potential ϪU↓ , in the turbed Landau spectrum, one can neglect the contribution case mϽ0 and the electron, bound by the potential ϪU↑ ,in ␦M(E) of the higher Landau levels, and the eigenfunctions Ͻ ␺ 10 the case m 0. i0␴ has the form In the case mϭ0 both carriers are captured by their cor- ␺ ␴ϭϪsin ␪␴␺ ␴ϩcos ␪␴␺ ␴ ͑A3͒ responding short-range potentials, and then their total cou- i0 ie b ␪ ϭ ⌬ pling strength, U↓ϪU↑ , determines the binding energy of with the mixing coefficient given by tan 2 ␴ 2Veb / ␴ , and ϭ ͉⌬ ͉ ␺ the exciton. Veb ͗e1 U b0͘. The wave functions b␴ describe the Landau state bound in a short range attractive potential,13

4. CONCLUSIONS ⌫͑␰͒ W␣ ͑␰͒ ␺ ͑␰͒ϭ ,0 ͑ ͒ b␴ ͓ ␲␺Ј͑␰͔͒1/2 ␰1/2 , A4 To conclude, we have found that a magnetic impurity lB 2 can bind a spin exciton in a 2DEG. It turned out that the ␰ϭ ␳ 2 ⌫ ␺Ј where 2 ( /lB) , , , and W␣,m are the gamma func- mechanism of the exciton capture is the spin-selective reso- tion, trigamma function, and Whittaker function, respec- nance scattering by the deep impurity levels. This spin se- tively. The index ␣ is determined by the corresponding lectivity stems from Hund’s rule in the particular case of ␣ϭ Ϫ1 Ϫ␧ 2 ␧ϭ ប2 eigenstate, 2 (1 b0␴lB) with 2m*E/ . When transition-metal impurities. The interaction in the second or- ␧ the level b0␴ is deep enough or the magnetic field is weak 2 der in scattering potential can be described in terms of an enough, i.e., ͉␧ ␴͉l ӷ1, the wave function ␺ ␴ ͑A4͒ has the ͓ ͑ ͔͒ b0 B b indirect spin exchange J see Eq. 14 , and only the longi- standard asymptotic form tudinal component of this exchange is essential for the for- Ϫဧ mation of a bound spin exciton. The transverse components e ␺ ␴ϳ . ͑A5͒ of this exchange give a contribution to the binding energy b ͱဧ only in fourth order in the scattering potential. ͉␣͉Ϸ 2 ␳2 ϭ␳ ␳ ␳Ϫ1ϭ␧ It is found that the spin-selective impurity potential is Here lB/2 b, P / b , and b b . Thus the local- ␳ always capable of binding the exciton in a state with the ization radii of the d electron ( d), bound Landau electron ␳ ␳ moment mϭ0 due to the edge van Hove singularity of the ( b), and free Landau electron (lB) obey the hierarchy d Ӷ␳ Ӷ density of states in a 2DEG. Excitons with m Þ0 can be also b lB . As a result one can safely neglect the contribution ␺ ͑ ͒ ͑ ͒ trapped, but then the conditions for the capture are more of the d component ie␴ A3 into the overlap integral 16 . ␤ ϭ ⌬ restrictive. Then in the case of weak scattering one has ␴ Veb / ␴ The analysis of the electronic structure of 3d transition Ӷ1(␪ Ӷ1), and the overlap integrals ͑16͒ can be approxi- metal impurities in GaAs-related systems shows that the mated by the following equations strongest binding potential is created by the light elements ϭ Ϫ␤2͒ ϭ Ϫ␤2͒ ͑ ͒ A p ͑1 ↑ J p↑ , B p ͑1 ↓ J p↓ . A6 ͑V, Cr͒. This work was supported by the German-Israeli Founda- Thus the spin dependence of the matrix elements ͑A6͒ is tion for Research and Development, Grant No. 0456- determined by the energy differences ⌬␴ in front of the in- 220.07195. K. K. thanks Israeli Science Foundation for sup- tegral and by the index ␣ of the Whittaker function. Having port ͑grant ‘‘Nonlinear Current Response of Multilevel in mind the difference in localization degrees of the wave Quantum Systems’’͒. The authors are indebted to Yu. Bych- functions ͑A4͒ and ͑19͒, we can estimate the magnitude of ϳ ␳ kov, S. Dickmann, T. Maniv, and I. Vagner for valuable dis- the overlap integrals J p␴ as J p␴ ( b␴ /lB), i.e., these inte- cussions. grals are sensitive both to the spin splitting and to the mag- netic field. The energy differences in the Hamiltonian ͑13͒ are determined by the short-range component of the impurity APPENDIX potential, which in our theory enters as a phenomenological ␻ In order to derive the impurity Hamiltonian for 2D elec- parameter 0 . In the weak-scattering limit one has Eb0␴ Ϫ ϭϪ␻ ϩ␤ ͑ trons in the Landau gauge, we use the fact that the resonance E0 0 ␴Veb if the short-range potential is scattering involves only the Landau states with mϭ0. These attractive͒.10 As a result, the latter factor is dominant in this states are included in the Hamiltonian ͑11͒, so one should limit, so that subtract them from Hi , which then acquires the form ϭϪ ␻ Ϫ␤ ͒ ͑ ͒ K␴ ͑ 0 ␴Veb . A7 ϭ ϩ ϩ ϩ ϩ Ϫ ϩ Ϫ ϩ Hi Eb0↑ ai0ai0 Eb0↓bi0bi0 E0↑a0 a0 Eb0↓b0 b0 . In accordance with Hund’s rule for 3d impurities ͑A1͒ ⌬↑Ͼ⌬↓ , and so K↑ϾK↓ . Then the bound Landau states with mϭ0 can be re-expanded Next we consider the situation where one of the reso- ␧ in free Landau states, nance d levels, namely ie↓ , is above the lowest Landau level, and the scattering is strong for the spin-down elec- ϭ ϭ ͑ ͒ ␪ ϳϪ␲ ␪ Ӷ ai0 ͚ A0nan0 , bi0 ͚ B0nbn0 A2 trons, ↓ /4. The inequality ↑ 1 still holds, since the n n estimates ͑A6͒ and ͑A7͒ for spin-up states are still valid. As ͑Ref. 13͒. As a result we come to Eq. ͑13͒. for the spin-down states, the coefficient sin ␪↓ is ϳ1/2 and In order to calculate the expansion coefficients A p , B p in 1 Eq. ͑15͒, one needs the wave functions of the bound electron K↓ϭϪ ͑␻ ϩV ͒. ͑A8͒ 2 0 eb in the lowest Landau level. The wave functions of the lowest localized Landau states correspond to the solutions Eb0␴ of The magnitudes of the scattering potentials U↑ and U↓ differ Eq. ͑9͒. If these states are deep enough below the unper- noticeably in this case. 806 Low Temp. Phys. 27 (99–10), September–October 2001 V. Fleurov and K. Kikoin

*E-mail: fl[email protected] and Superlattices, edited by D. J. Lockwood and J. F. Young, Plenum, New York ͑1991͒. 9 1 K. A. Kikoin and V. N. Fleurov, Transition Metal Impurities in Semicon- R. J. Nicolas, R. J. Haug, K. von Klitzing, and G. Weimann, Phys. Rev. B ductors, World Scientific, Singapore ͑1994͒, Chapt. 8. ͑ ͒ 37,1294 1988 . 10 P. Dahan, V. Fleurov, K. Kikoin, and I. Vagner, cond-mat/0106339 ͑sub- 2 ´ ͑ ͓͒ I. V. Lerner and Yu. E. Lozovik, Zh. Eksp. Teor. Fiz. 78, 1167 1980 Sov. mitted to Phys. Rev. B͒. Phys. JETP 51,588͑1980͔͒. 11 ͑ ͒ 3 G. Picoli, A. Chomette, and M. Lannoo, Phys. Rev. B 30, 7138 1984 . Yu. A. Bychkov, S. V. Iordanskii, and G. M. Eliashberg, JETP Lett. 33, 12 V. N. Fleurov and K. A. Kikoin, J. Phys. C 19,887͑1986͒. ͑ ͒ 143 1981 . 13 Y. Avishai, M. Ya. Azbel, and S. A. Gredeskul, Phys. Rev. B 48, 17280 4 ͑ ͒ C. Kallin and B. I. Halperin, Phys. Rev. B 30, 5655 1984 . ͑1993͒. 5 ͑ ͒ C. Kallin and B. I. Halperin, Phys. Rev. B 31, 3635 1985 . 14 Yu. A. Bychkov, T. Maniv, I. D. Vagner, and P. Wyder, Phys. Rev. Lett. 73, 6 ͑ ͒ D. Antoniou and A. H. MacDonald, Phys. Rev. B 43, 11686 1991 . 2911 ͑1994͒. 7 J. P. Longo and C. Kallin, Phys. Rev. B 47, 4429 ͑1993͒. 8 A. Pinczuk, D. Heiman, S. Schmitt-Rink, C. Kallin, B. S. Dennis, L. N. This article was published in English in the original Russian journal. Repro- Pfeiffer, and K. W. West, in Light Scattering in Semiconductor Structures duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Shell effects in alkali metal nanowires A. I. Yanson* and J. M. van Ruitenbeek

Kamerlingh Onnes Laboratorium, Leiden University, PO Box 9504, NL-2300 RA Leiden, The Netherlands I. K. Yanson

Kamerlingh Onnes Laboratorium, Leiden University, PO Box 9504, NL-2300 RA Leiden, The Netherlands, B. Verkin Institute for Low Temperature Physics and Engineering of the National Academy of Sciences, 47 Lenin ave., 61103, Kharkov, Ukraine ͑Submitted July 3, 2001͒ Fiz. Nizk. Temp. 27, 1092–1109 ͑September–October 2001͒ After making a cold weld by pressing two clean metal surfaces together and then gradually separating the two pieces, a metallic nanowire is formed, which progressively thins down to a single atom before the contact is lost. We show evidence that the radii of such nanowires are influenced by electronic shell filling effects, in analogy to electronic shell structure in metal clusters. A modulation of the shell effect, known as supershell structure, is clearly observed. For sodium and potassium at larger diameters there is a crossover to crystalline facets with shell- closings corresponding to the completion of an additional atomic layer. We conclude that a complete analogy exists between shell structure in metallic clusters and nanowires. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1414569͔

1. INTRODUCTION these objects electrons are not confined along the wire axis. In spite of this, shell effects manifest themselves as peaks in The understanding of processes that occur on an atomic a histogram for the electrical conductance of the wire re- scale when two materials are brought into contact is of fun- corded for wires while the wire diameters are changed, damental importance to problems such as adhesion, hard- showing the most stable configurations. Since the conduc- ness, plastic and elastic response to external strength, frac- tance depends on the wire cross-sectional area, the latter 1 ture, friction, and wear. For conducting materials a new quantity gives, on the one hand, the number of conduction phenomenon appears due to quantization of conduction electron modes in the two lateral dimensions, and, on the 2–4 channels. Novel techniques enable one to create and study other hand, the number of atoms. Both of these effects, elec- the properties of nanowires down to a chain of single tronic and atomic, are distinctly seen in the conductance his- 5,6 atoms. For these experiments, as well as for the above- tograms, which exhibit their coexistence and crossover. mentioned problems, the processes connected with stability Thus, there appears to be a complete analogy between the and cohesive energy of nanoscopic contacts are of primary shell effects in metallic clusters and nanowires. importance. The stability of microscopic and nanoscopic objects is 2. EXPERIMENTAL TECHNIQUE tightly connected with the so-called shell effects. In three dimensions, shell effects are concerned with the numbers of For our study of nanowires we use ͑Fig. 1a͒ the me- microscopic particles ͑baryons in nuclei, electrons in an chanically controllable break-junction ͑MCB͒ technique de- atom, atoms in a cluster͒ for which an enhanced stability of veloped in Ref. 14. The metallic sample was fixed by Stycast the system is found. These are called ‘‘magic numbers.’’ For epoxy on a flexible substrate made from phosphor bronze the periodic table of elements the noble gases, with fully ͑bending beam͒ and supported by three points: two are on occupied electronic shells, acquire the highest stability, since the upper side of the bending beam ͑counter supports͒, and they possess the largest ionization potential and are chemi- the third is at the opposite side supported by a movable piezo cally inert. For the clusters emitted by a vapor jet in a mass driver. A metallic sample, prepared as a thick wire, is elec- spectrometer, the abundance of ‘‘magic masses’’ is enhanced, trically isolated from the bronze substrate. The wire was cir- since these clusters posses higher stability as compared to the cularly notched in the middle in order to concentrate the others.7–9 pulling strength at the center while the substrate was bent by In metallic clusters, one distinguishes two different shell mechanical ͑not shown͒ and piezo drivers. By fixing the effects: one is connected with conduction electrons, for sample as close as possible near the notch, a large reduction which the cluster behaves like a giant atom,7–9 and the other ͑100–1000͒ is achieved in the ratio between movement of relates to the configuration of atoms in the cluster.9,10 Both the electrodes with respect to each other and the expansion of are due to minima in the cluster free energy, but the first the piezo driver, in the perpendicular direction. requires conduction electrons, while the second also applies In order to study alkali metals, two important modifica- to nonconducting materials. tions are made. First, due to the extraordinary reactivity of In this paper we describe in detail the shell effects dis- alkaline elements, contact with the ambient environment covered by us in an open system: a metallic nanowire.11–13 In must be avoided. This is achieved by making all the prepa-

1063-777X/2001/27(9–10)/14/$20.00807 © 2001 American Institute of Physics 808 Low Temp. Phys. 27 (99–10), September–October 2001 Yanson et al.

FIG. 2. Normalized conductance histograms of Ͼ1000 individual conduc- tance traces of Na, K, and Li measured at 4.2 K under cryogenic vacuum at a constant voltage bias of 10 mV. The curves for K and Na are shifted for clarity by 0.15 and 0.3, respectively. Typical individual conductance traces ͑ ͒ ϭ ϭ 2 scans are shown in the inset. g G/G0 , where G0 2e /h, is the reduced conductance.

wire at large conductances, we prefer to use the smallest possible biases. Our experience leads to the conclusion that up to a bias of ϳ10 mV the heating can be neglected. The drift and calibration of the current-to-voltage converter is verified against standard conductors of 1, 10, and 100G0 , FIG. 1. a—Schematic drawing of a substrate with a sample mounted in a ensuring an overall accuracy in the conductance better than Ͼ three-point bending configuration. The bending of the substrate due to the 1% for G 10G0 . Scans are taken continuously by ramping movement ͑or expansion͒ of the piezo over a distance ⌬, shown in gray, the displacement ␦ϵl of the electrodes with respect to each ␦ causes the epoxy droplets to move apart by , leading to the breaking of the other, using the piezo driver. Each individual curve of the junction. The amount of bending is exaggerated for clarity. b—A model of an alkali break junction. The principle is the same here as in part ͑a͒ but the conductance versus the displacement, G(l), is recorded in two epoxy drops are replaced by four M1 bolts with nuts fixing the sample ϳ0.1 seconds from the highest conductance into the tunnel- to the substrate. The length of the substrate remains 22 mm. ing regime. The curves are digitized with 16-bit resolution and a data acquisition rate of ϳ105 point/s. The conductance scale is divided into 1500 bins and data points from up to rations in paraffin oil. The sample is cut into the shape of a ϳ105 individual scans are accumulated automatically, creat- rectangular bar, at the center of which we cut a notch. Since ing a conductance histogram at the given experimental con- the bar with the notch is also covered by the oil, it cannot be ditions ͑temperature, voltage bias, indentation depth͒. For glued and must be fixed mechanically. For this purpose we most cases we have found that about 103 scans are enough to use four tiny bolts and nuts that are electrically isolated from achieve convergence in the shape of the histograms. The the substrate ͑Fig. 1b͒.4 After mounting, the whole setup is speed of scanning is limited from above by the data acquisi- rapidly placed in an evacuated metallic can and put into a tion rate of the electronics, and from below by a reasonable cryostat with an environmental temperature of 4.2 K. The time required for the experiment ͑normally, several hours͒. walls of the can that are being kept at helium temperature To compare different histograms, we normalized them by serve as a cryogenic pump ensuring a very high vacuum their area over a fixed conductance range. around the sample. The bar is then broken at the notch by mechanical bending of the substrate. The cyclic electrical movement of the piezo drive rod periodically sweeps the 3. RESULTS electrical conductance between the electrodes from a desired value ͑of the order of tens to hundreds of conductance 3.1. Experiment quanta1͒͒ down to complete separation ͑conductance much Low-temperature histograms and scans. In Fig. 2 we less than one quantum unit͒. The important advantage of the show histograms for Na, K, and Li taken at Tϭ4.2 K, along MCB device is that the cryopumping action of the low- with typical scans ͑see inset͒. All histograms display the temperature environment ensures that the freshly exposed characteristic series of peaks in succession at 1, 3, 5, 6, in fracture surfaces are not polluted by adsorbates. units of reduced conductance2͒ which has been shown in the- The conductance is measured in a four-terminal dc con- oretical and experimental works4,15,16 to be the hallmark of stant voltage mode. A fixed bias voltage is applied, in the quantum mode degeneracy for electronic wave functions in a range 1–100 mV, so that the current through the contact is waveguide with circular cross section. This degeneracy proportional to its conductance. Since the high bias heats the arises from the azimuthal quantum number, which takes two Low Temp. Phys. 27 (99–10), September–October 2001 Yanson et al. 809 values ͑Ϫ and ϩ͒. When this happens ͑at gϭ1, 3, 6, and so on͒ the conductance increases by 2 quanta for increasing nanowire diameter. First we discuss the positions of these peaks. They are more or less shifted to the left compared to the integer val- ues. The least shift occurs for the gϭ1 peak and it increases noticeably for higher g. Phenomenologically, this increase can be described by a series resistance of the order of ϳ100⍀, originating from the backscattering of electron wave functions by defects and surface irregularities,4,17,18 and inelastic scattering by phonons. The latter may be the cause for a noticeably larger shift for Li, for which it is known that the electron–phonon interaction parameter is several times larger than for Na and K.19 The intensities of the first few peaks for Na and K, on the one hand, and Li on the other hand, are also quite differ- ent. The height of the peaks at gϭ1 and 3 in Li is much lower than for Na and K. On the contrary, the peaks at g ϭ5, 6 in Li is noticeably higher than those for Na and K. This may be explained qualitatively by a much higher mo- bility of Li atoms even at helium temperature, which may be due to tunneling because of the light Li mass and the de- crease in the potential barriers at the surface, making a con- tact of just a few atoms unstable. Next, we draw attention to the broad maxima at gϷ8, 11, 15, which are slightly more intense in Li, presumably again because of the enhanced mobility of Li surface atoms. We shall see below that these broad peaks are precursors of the electronic shell effect for alkali metal nanowires. To- FIG. 3. a—Typical individual scans for Li break-junctions at two different gether with the peak at gϭ5 they give the most favorable temperatures. 1 piezo volt corresponds approximately to 0.5 Å. b—Long diameters, corresponding to the local minima in the elec- plateau at 140 K, shown on a logarithmic scale in the inset. tronic part of the thermodynamic potential. As to the peak at gϭ5, this point of view is supported by the fact that, accord- shell structure of the nanowires. At low temperatures, the 20 ing to the theoretical calculation of the sodium histogram, piezo voltage versus elongation of the neck can be calibrated the intensity of the gϭ5 peak should be noticeably smaller using the Fowler-Nordheim formula or the Gundlach than the peak at gϭ6. In fact, in the experimental histograms oscillations22 for the tunnel regime at biases of the order of a ͑see Fig. 2͒ they are approximately of the same intensity. few volts. Another confirmation that the gϭ5 peak is not connected By storing several intermediate histograms during the with the fully open conducting channel comes from Ref. 18, accumulation of the final histograms, for the full number of where its behavior is radically different from that of the scans for a given contact, one can obtain an impression of peaks at gϭ1, 3, 6. Thus, we propose that the enhanced their evolution with increasing numbers of scans. One can intensity of the gϭ5 peak is due to the broad maximum of observe both the purification of the metal under the contact, the favorable diameters dictated by the electronic shell ef- which leads to more distinct shell-effect structure with in- fect. creasing number of indentations, as well as to the opposite Scans taken at helium temperature and shown in the in- behavior, which we ascribe to the adsorbed impurities.23 set in Fig. 2 exhibit smooth variations ͑plateaus͒ interrupted Temperature dependence of the scans. Let us describe by sharp decreases ͑steps͒, evidencing that the elastic elon- the temperature dependence of the scans, taking Li as ex- gations are followed by atomic rearrangements. This points ample ͑Fig. 3͒. Figure 3a shows two scans for low and el- at the crystalline structure of the neck. Only for the low-lying evated temperatures, typical for all the alkali metals studied. conductances do the plateaus coincide with quantum units As the temperature increases, the low-lying conductance pla- ͑Ϸ1, 3, 5, 6͒, which can be even seen statistically on the teaus become unstable. Moreover, the piezo voltage is no average scan plotted by integrating the histogram. For higher longer the only parameter controlling the evolution of the conductance, gտ7, though the steps in the scans look as neck, as a result of the strongly enhanced diffusion of the sharp as for lower g, no peaks on the low-temperature histo- surface atoms. The atoms in the wire experience tension not gram appear. This means that the plateaus occur with nearly only from the external force but also from the Laplace pres- equal probabilities for any conductances in different scans, sure. This means that upon putting the external force to zero leading on average to the smooth background in the histo- the conductivity of the neck still evolves with time until the grams. According to Ref. 21, from the smooth part of the contact disappears. scans one can find the approximate shape of the neck. We We mentioned above that Li atoms revealed increased shall turn to this issue below when discussing the atomic diffusion when we considered the low intensity of the gϭ1 810 Low Temp. Phys. 27 (99–10), September–October 2001 Yanson et al. peak in the Li histograms at helium temperatures. The en- hanced mobility of the atoms also leads to the anomalous behavior of the conductance trace G(Vpiezo) at high tempera- tures shown in Fig. 3b, where it steeply drops to zero from a fairly high conductance. For Li at Tϭ140 K all steps on the G(Vpiezo) scans disappear and all structure in the histograms disappears except at gϷ5. This peak in the histogram be- comes very high, corresponding to anomalously long pla- teaus in the scans. The length of the plateaus reaches several tens of Å, while typical plateaus correspond to less than 1 Å. The same qualitative behavior was observed for Na, although for this metal it was more difficult to register scans at such elevated temperature. We interpret this phenomenon in the following way. At TϷ140 K the Li nanowire ‘‘melts,’’ although the tem- perature of the bulk melting point is much higher. By this we mean that the diffusion of atoms over the length of the nanowire becomes much faster than our data acquisition rate. According to Refs. 24 and 25, a strong depression of the melting point is expected for small systems due to the Laplace tension. The change of the melting temperature Tm compared to the bulk value T0 is approximately given by Ϫ ϭ ␴ Ϫ␴ ␴ (Tm T0)/T0 2( 1 2)/QR, where 1,2 is the surface tension in the liquid and solid phases, respectively, Q is the heat of fusion, and R is the wire radius.26 Thus we suggest that at the starting point of the long plateau a liquidlike me- tallic neck appears, which connects the solid banks. Further elongation of the neck proceeds continuously, decreasing its diameter as the wire is stretched, approximately as a power FIG. 4. Temperature evolution of conductance histograms for Na and Li. ϰ n ӍϪ The histograms have been normalized to the total area under the curves. law g (Vpiezo) , with n 1.3. Temperature dependence of the histograms. The tem- perature dependences of Na and Li histograms are shown in oscillations can be obtained from the Fourier transform of Fig. 4. One recognizes the low-temperature curves similar to the experimental histograms. those shown in Fig. 2. While raising the temperature, the low-conductance peaks gradually disappear due to increasing thermal instabilities. On the other hand, for thicker wires a 3.2. Electronic shell and supershell effects new succession of strong peaks grows, owing to the in- In this Section we describe the observation of electronic creased mobility of atoms at elevated temperatures. We shall shell-effect oscillations11,12 in alkali metal nanowires. The consider these series of peaks in detail in the next Sections. experimental results depend on several factors discussed Here we only point out that the positions of these peaks on above. We emphasize once again that all shell-effect struc- the conductance axis does not correspond to those that might tures are observed at elevated temperature, at which atoms arise due to individual quantum modes, which would incre- have enough mobility to occupy the most favorable positions ment either by one or two conductance quanta. Instead, these in the lattice structure. This structure corresponds to the local peaks are broader and reveal a separation that grows with minima of the total free energy of the nanowire. conductance. The intensities of these peaks depend, among Histograms and peak positions. A typical histogram for other parameters, on the depth of indentation, which can ap- sodium is shown in Fig. 5a. The positions of the main peaks, proximately be characterized by the highest value of the con- labeled by sequential numbers, are regularly spaced as a ductance scale in the histogram. A second parameter is the function of g1/2. That means that they periodically appear as voltage bias at which the scans are recorded. Unfortunately, a function of the radius of the wire, since for a circular there are additional parameters that we can hardly control nanowire the radius of the smallest cross section R, obtained with the technique used. These include possible surface con- from the Weyl semiclassical27 expression for a ballistic re- taminations and the orientation of the crystallites on the two gime, reads: electrodes. We stress that these drawbacks mainly concern k R 2 2 k R 2 the intensity of the peaks and much less their positions on the Ϸͩ F ͪ ͩ Ϫ ͪ Ϸͩ F ͪ ͑ ͒ gW 1 , 1 conductance axis, while the period of the peak series remains 2 kFR 2 2ϭ almost unaltered. Since, as will be shown below, there are where kF is the Fermi wave vector and (kFR/2) gS is the two mechanisms leading to different series of peaks with Sharvin conductance. We prefer to plot g1/2 along the ab- different periods ͑namely, electronic and atomic shell series͒, scissa, since it is the experimentally measured quantity, while the positions of the peaks depend upon the relative intensity the inferred radius depends somewhat on the shape of of these mechanisms. Fortunately, the true frequencies of the nanowire. For example, for a contact in the form of an orifice Low Temp. Phys. 27 (99–10), September–October 2001 Yanson et al. 811

FIG. 6. a—Conductance histogram for sodium from Fig. 5a with a smooth FIG. 5. Conductance histogram for sodium, constructed from Ϸ14000 indi- background subtracted ͑the original curve and the background are shown in vidual traces recorded at Tϭ90 K and at 100 mV voltage bias ͑a͒; conduc- the inset͒. The arrows point at the positions of the minima in the oscillation tance histogram for potassium, constructed from ϳ5000 individual traces amplitudes. b—Fourier spectrum of the curve in part ͑a͒. recorded at Tϭ100 K and 8 mV voltage bias ͑b͒; conductance histogram for lithium, constructed from ϳ8000 individual traces recorded at Tϭ100 K and 10 mV voltage bias ͑c͒. We count the split peak #3 as a single peak, as missing, just as for Na. The periods are close to what is for Na ͑see ͑a͒͒. The peaks are numbered as shown, and a plot of peak positions vs their index number is fit by a straight line in the insets. The observed in Na, although a small change in the slope is often slopes are 0.581Ϯ0.005, 0.54Ϯ0.005, and 0.62Ϯ0.02 for Na, K, and Li, observed for different samples. The modulation of the peak respectively. Note that the horizontal axis is given as the square root of g. heights also varies from contact to contact. While a modula- tion is clearly seen in the Li histogram ͑Fig. 5c͒, for the given K contact ͑Fig. 5b͒ only a nonmonotonic behavior Ӎ Ϫ one has gW gS͓1 1/(kFR)͔, which makes the correction around the peaks with indices 3–5 is found. In the next Sec- to the Sharvin conductance smaller.16 The inset in Fig. 5a tion we present a different K histogram, which exhibits much demonstrates that the peak positions are periodic, with a pe- more modulation by the superposition of different shell ef- riod of ϳ0.58. We see that at elevated temperatures the g fects ͑Fig. 7͒. ϭ1 peak is absent and the highest intensity is reached at g Fourier transformations. Determination of the average Ϸ16. The shape of the background is determined by the period from the slope of the peak positions in terms of g1/2 global variation of the conductance with elongation, which in versus their indices ͑see insets in Fig. 5͒ gives limited infor- turn depends on the effective length of nanowire that de- mation about the probabilities for observing specific conduc- forms during elongation21,28 ͑see below͒. This length is lo- tance values. For example, in this way, one cannot account cated near the narrowest cross section of the nanowire and for the nonmonotonic behavior of the peak heights. The can be qualitatively described with the model considered in problem arises, whether to count the peaks with smaller am- Ref. 21. The latter is valid when diffusion of the atoms from plitude. the neck to the electrodes can be neglected.3͒ Superimposed Fortunately, we can make use of a direct Fourier analysis in the main periodic structure we note an oscillating behavior of the histograms, which exposes a superposition of several of the envelope of the maxima in the histogram of Fig. 5a. periodic components. A typical example for sodium is shown This modulation is the signature of the supershell effect and in Fig. 6. In order to avoid a huge low-frequency tail in the will be considered in more detail in the section on the Fou- Fourier spectra, we subtract a smooth background shown by rier transformation. Finally, we draw attention to the low- the dashed curve in the inset of Fig. 6a. The resulting curve intensity oscillations at g1/2տ7.5. They have a much smaller is displayed on the main panel of Fig. 6a. It reveals a clear period and are due to the crossover from the electronic to the low-frequency modulation with a period of ϳ4.0. Here, we ͑ ͒ 1/2 atomic shell effect see below . use the variable kFR instead of g for the abscissa, by use Similar periodic structure is seen for potassium ͑Fig. 5b͒ of the Weyl expression ͑1͒, which will prove convenient for and lithium ͑Fig. 5c͒. The periods shown in the insets are comparison with the theoretical results in the discussion sec- 0.54 and 0.62 for K and Li, respectively. Maximal ampli- tion below. The Fourier spectrum for this histogram is shown tudes are observed at g1/2Ӎ4 – 5, and the peak at gϭ1is in Fig. 6b. It contains two prominent frequencies, 0.64 and 812 Low Temp. Phys. 27 (99–10), September–October 2001 Yanson et al.

TABLE I. First few shortest classical periodic orbits and their frequencies in circular geometry. The circle (tϭ1,pϭϱ) marks the end of the first band (tϭ1); the second one contains trajectories which make two revolutions around the center (tϭ2), the third—three. etc.

FIG. 7. a—Conductance histogram for potassium at Tϭ50 K, constructed from 1000 individual scans recorded at 125 mV voltage bias, after subtract- ing a smooth background ͑the original curve and the background are shown in the inset͒. b—Fourier spectrum of the curve in part ͑a͒.

Finally, let us consider the Fourier transform of the Li histogram from Fig. 5c. Figure 8b shows three peaks of

0.89, where the first has a lower intensity than the second. The latter of the two frequencies is very close to the ͑aver- age͒ one obtained for the magic numbers in spherical clusters.11 As will be discussed below, for clusters the inten- sity of the first principal frequency is negligibly small, and the observed frequency equals the average of triangular and square orbits: (0.83ϩ0.90)/2ϭ0.86 ͑see Table I͒. The pres- ence of the lower-frequency component for the nanowires explains why the frequency of the second prominent peak is a little bit higher than what would be obtained from the averaged period ⌬(g1/2)Ӎ0.58 given by the slope in the inset of Fig. 5a.4͒ In Fig. 6b the low-intensity peaks at frequencies above 1 are presumably due to the harmonics and combina- tions of the two principal frequencies. Considering the histogram for potassium ͑Fig. 7a͒, which is different from the one in Fig. 5b, we see that there are again two prominent peaks in the Fourier transform, at 0.63 and 0.84, close to those observed for sodium. In addi- tion, there is an appreciable maximum around 2.2, which is difficult to explain in terms of overtones but stems from the atomic shell effect, as we will see below. For the given K ⌬ Ӎ histogram one obtains an average period of (kFR) 1.15 from the slope of the peak positions, giving an average fre- quency of 1/⌬(k R)ϭ0.87. This nearly coincides with the F FIG. 8. a—Conductance histogram for lithium from Fig. 5c with a smooth upper principal frequency because of the small intensity of background subtracted ͑the original curve and the background are shown in the first principal frequency. the inset͒. b—Fourier spectrum of the curve in part ͑a͒. Low Temp. Phys. 27 (99–10), September–October 2001 Yanson et al. 813

FIG. 10. Sodium histogram constructed from 10424 individual conductance ϭ FIG. 9. Potassium histogram constructed from 1000 individual conductance scans recorded at 100 mV voltage bias and at T 100 K. The peaks are traces recorded at 1 mV voltage bias and at Tϭ100 K. The peaks are num- numbered sequentially and their position vs number is plotted in the inset. A bered sequentially and their position is plotted as a function of this number straight-line fits give slopes of 0.62 and 0.224 for the electronic and atomic in the inset. Straight solid lines show the slopes 1.447 and 0.241 for full series, respectively. One of the typical histograms for potassium is shown ͑ ͒ atomic shells and single hexagonal facets, respectively. A linear fit for the for comparison dotted curve . electronic-shell points gives 0.61 ͑dashed straight line͒. A linear fit to the experimental data starting with number 6 gives a slope of 0.224. Along with 1/2 g values on the horizontal axis the corresponding kFR values are given on top, according to formula ͑1͒. can see that the transition between these two sets of oscilla- tions ͑electronic and atomic͒ is quite sharp, although the shorter period pattern extends a little with reduced amplitude about the same intensity. Those at frequencies 0.66 and 0.93 into the lower conductance range, where it overlaps with the are of the same origin as the principal frequencies in Na and electronic shell oscillations. K. As to the lowest frequency at 0.25, it may be due to an For sodium, the crossover is found at larger diameters anomalously strong interference of the principal frequencies, than for potassium ͑Fig. 10͒ with periods equal to (0.62 giving rise to a difference frequency with a period of 4.2. Ϯ0.05)g1/2 and (0.224Ϯ0.001)g1/2 for the electronic and This may be due to the fact that the intensity of the first atomic shells, respectively. For comparison in Fig. 10 we principal peak ͑at 0.66͒ is as strong as the second one. It is plot the histogram for the potassium nanowires as a thinner interesting to note that both principal Fourier frequencies in curve. Some discrepancies between the Na and K histograms Li are higher than in Na and K. However, the average fre- are seen in the low conductance range, which are mostly quency of (0.66ϩ0.93)/2ϭ0.78 is lower than in Na, corre- influenced by the different voltage bias. However, the behav- sponding to a larger average period ⌬(g1/2)ϭ0.64 ͑or ior remains basically the same. ⌬(k R)Ӎ1.27͒. Indeed, since in Li the intensities of the two F Superposition of electronic and atomic-shell structure at principal peaks are approximately the same, we can simply low conductances. Atomic-shell oscillations can be ob- average their frequencies, while this is would not be correct served at conductances starting as low as gϷ6 – 7. In Fig. 11 for Na and K. we show some of the histograms recorded in a single series under seemingly the same experimental conditions. In the 3.3. Atomic shell effect13 first recording no distinct oscillations were observed ͑not Searching for the shell oscillations at still higher conduc- shown͒. Each trial, which consisted of many scans, was per- tance, we have found a new series that appears to be due to formed by making a contact with a resistance of about 1 ⍀ a purely-classical effect, similar to that found in clusters. As (ϳ104g) and then mechanically receding the electrodes to in clusters, we named it an ‘‘atomic shell effect’’ and discuss bring the conductance down to the tunneling regime. The its origin in the next Section. We start by presenting the main next several trials are displayed in Fig. 11 ͑curves 1–5͒.Itis experimental facts concerning this effect. seen that for this particular sample the histograms change Crossover from electronic to atomic shell effect. The drastically from those resembling the atomic-shell oscilla- most spectacular results are achieved for potassium. The cor- tions ͑curves 1, 2, and 3͒ to ones having the electronic shell responding histogram is shown in Fig. 9. It shows two types periodicity ͑curves 4 and 5͒. Presumably, this is due to the of oscillations with a relatively sharp crossover at g1/2Ϸ6. In increase in the mobility of the surface atoms. Namely, for the the inset, the reduced conductances, corresponding to the last trials the mobility of atoms at the surface is increased so peaks in the histogram, are plotted against their sequentially much that no crystalline facets ͑see below͒ could be created numbered indices. The index number of the peaks is also during the acquisition time for the scans. Other causes may shown in the histogram, where they are labeled at increments be the random crystal orientation of the neck, creating a fa- of five. The first few points in the inset, corresponding to the vorable condition for one of the two shell effects, and alter- strong peaks in the histogram, have an approximate slope of ing of the shape of the nanowire during many indentations. 0.61, which agrees reasonably well with the slope deter- Interestingly, the extreme shape of histograms presented in mined for the electronic shells. Beyond the first few points, Fig. 11 ͑curves 4 and 5͒ reproduces quite satisfactorily the up to 48 points obey a linear relation with a slope of 0.225 electronic-shell histograms presented above. The same is true Ϯ0.001. Structure in the histograms with this particular pe- for the atomic-shell histograms at low conductances ͑curves riod has been reproducibly observed for many contacts. One 2 and 3͒. As an illustration, in Fig. 12a we show by a thick 814 Low Temp. Phys. 27 (99–10), September–October 2001 Yanson et al.

FIG. 13. The averaged period of the oscillations in Fig. 9 plotted vs peak index. The vertical dotted lines mark the positions of peaks with maximal amplitude ͑numbers 7, 13, 19, 26͒. The gray staircase is a guide to the eye to identify the step structure. A fragment of a conductance histogram is shown ⌬ ϭ Ϫ in the inset to explain the averaging procedure ͗ gi͘ 1/2͓(gi giϪ1) ϩ Ϫ (giϩ1 gi)͔. An artist’s impression of the atomic positions in the normal cross section of the wire is sketched in the lower right corner. For simplicity, we have drawn a two-dimensional structure. The six facets symbolize the six FIG. 11. Evolution of shell-oscillations obtained in successive runs in the data points at each plateau, while the height of the step corresponds to ϭ ϭ same experiment for a potassium nanowire at T 80 K,Vbias 10 mV. In successive layers for a close-packed structure. curves 2 and 3 the atomic-like period is clearly seen, while for curves 4 and 5 the period is due to the electronic shell effect. Each histogram consists of from several thousands up to several tens of thousands of scans.

reproducibility of the peak positions is quite good, despite ͑ line the histogram for another sample recorded in an experi- important variations in the amplitudes. One should not think ment approximately a year later͒ and superimposed on curve that the electronic-shell structure is negligibly small for these 2 from Fig. 11 ͑note the change in the x coordinate͒. The curves. The Fourier transformation shown in Fig. 12b for the solid curve in Fig. 12a clearly shows the characteristic fre- quencies for both shell effects: electronic peaks at 1.2 and 1.8 and the atomic peak at 4.4. Other features with lower intensity may be explained as harmonics and sub-harmonics of the principal frequencies. Additional structure of atomic-shell oscillations. Occa- sionally, the atomic-shell oscillations reveal additional struc- ture in their amplitude and periodicity. The peaks can be grouped by sixes, based on the amplitudes, as seen for ex- ample in Fig. 9. Here, the thin vertical lines at g1/2Ϸ6.3, 7.6, 8.9, and 10.5 indicate higher amplitudes and correspond to the index numbers 7, 13, 19, and 26. For other contacts, there may be other envelope patterns. The period of these higher- amplitude peaks equals 1.36Ϸ0.224ϫ6, which is very close to a single close-packed monatomic layer covering the nanowire ͑see below͒. Moreover, the period between adja- cent peaks as a function of the reduced conductance g ap- pears not to scatter randomly around a linear dependence as one might expect. A linear dependence follows when the period is constant as a function of g1/2. Instead, when the distance between the adjacent peaks is averaged over 3 points ͑as shown in the inset of Fig. 13͒ in order to smoothen fluctuations, they can again be grouped by approximately six FIG. 12. a—Two conductance histograms for potassium showing mainly peaks ͑Fig. 13͒. For this particular contact, the averaged pe- atomic-shell oscillations. The dotted curve is curve 2 from Fig. 11, while the full curve was taken almost a year later. Good reproducibility is seen for the riod exhibits a smooth staircase in which the vertical steps measurements. b—Fourier spectrum of the full curve in part ͑a͒. Superpo- closely coincide with the higher-amplitude peak indices sition of two shell effects is seen from the peaks in the Fourier spectrum at ͑numbers 7, 13, 19, and 26͒. The difference in height of the 1.2 and 1.8, and 4.4 for electronic and atomic principal frequencies, respec- ϭ ϭ plateaus is approximately one G . This behavior will be ex- tively. T 80 K, Vbias 10 and 20 mV for the dotted and solid curves, re- 0 spectively. plained in the next Section. Low Temp. Phys. 27 (99–10), September–October 2001 Yanson et al. 815

4. DISCUSSION 4 2m 1/2 ⍀ϭ Ϫ␮ ϭϪ ͩ e ͪ ͑␧ Ϫ␧ ͒3/2 E N L 2 2 ͚ F mn , 4.1. Electronic shells 3 ␲ ប m,n Exact quantum mechanical result. To understand semi- whose derivative with respect to the length gives us the force quantitatively the electronic shell effect in the formation of ⍀ 1/2 d 2me metallic nanowires, we use the simple model of a metallic FϭϪ ϭͩ ͪ circular cylinder15,29,30,33 connecting two metallic reservoirs dL ␲2ប2 ͑electrodes͒. A more realistic shape for the wires21 will be 4 considered later. When the wire diameter d is comparable ϫ ͩ ͑␧ Ϫ␧ ͒3/2Ϫ ␧ ͑␧ Ϫ␧ ͒1/2ͪ ͚ F mn 2 mn F mn . with the de Broglie wave length of conduction electrons at m,n 3 the Fermi energy, quantum size effects in the lateral dimen- The density of states, the thermodynamic potential, and the sions lead to a finite number of conductance channels, each force are calculated as functions of the wire radius for a corresponding to an electron wave propagating along the ␧ given Fermi energy using the formulas wire axis z. In order to obtain the energy spectrum (kz)of these modes, one solves the Schro¨dinger equation to obtain 1 ␥ 2 Ϫ1/2 ͑ ͒ϭ ͫ Ϫͩ mn ͪ ͬ D kFR ␲␧ ͚ 1 , ប2 2 F m,n kFR kz ␧͑k ͒ϭ␧ ϩ , ͑2͒ z mn 2m 2ប2k5 V 1 ␥ 2 3/2 e ⍀͑ ͒ϭ F ͫ Ϫͩ mn ͪ ͬ ͑ ͒ kFR ␲2 ͑ ͒2 ͚ 1 , 5 where 3 me kFR m,n kFR

2 ប2 3 ␥ 2 3/2 ប kF 2 mn ␧ ϭ ␥2 F͑k R͒ϭ ϫ͚ ͫ ͩ 1Ϫͩ ͪ ͪ mn 2 mn, F ␲m 3 k R 2meR e m,n F m is the electron mass, ␥ is the nth zero of the mth Bessel ␥2 ␥ 2 1/2 e mn Ϫ mn ͩ Ϫͩ mn ͪ ͪ ͬ ͒2 1 . function of the first kind, R is the radius of the nanowire, d ͑kFR kFR ϭ 2R, and kz is the z component of the electron wave num- The results for a sodium nanowire with a constant vol- ber. In any single one-dimensional subband the total number 3 of states per unit length of the wire ͑including spin degen- ume of 10 nm and the bulk Fermi energy are plotted in Fig. eracy͒ is 14. Whenever the bottom of another subband falls below the Fermi energy, the density of states and the force exhibit sin- ͉ ␧͉͒ 1/2 2 kz͑ 2me gularities. The force, being the derivative of the thermody- N ͑␧͒ϭ2 ϭ2ͩ ͪ ͱ␧Ϫ␧ , ͑3͒ mn 2␲ ␲2ប2 mn namic potential with respect to the length of the cylinder, describes the response of the system to elongation. When its and the density of states is value is nonzero the wire tends to shorten ͑negative force͒ or stretch ͑positive force͒ spontaneously in order to reduce its ͑␧͒ 1/2 dNmn 2me 1 free energy; only when the force is zero is a true state of D ͑␧Ϫ␧ ͒ϵ ͩ ͪ . ͑4͒ mn mn d␧ ␲2ប2 ͱ␧Ϫ␧ stability achieved. We should also take into account that in mn the classical limit of large diameters we should recover a Let us now consider a cylinder of finite length L, radius R, constant destabilizing force resulting from the surface ten- and volume V, which is elongated at the ends so that its sion. Therefore we do not have truly stable nanowires but volume Vϭ␲R2L remains constant. To use expressions ͑2͒– rather ones with enhanced stability at the radii where the ͑4͒ we should impose the condition LӷR, which allows us force is minimal. Those minima in the absolute value of the to consider the cylinder as quasi-infinite in the z direction. force give the diameters at which the wires are particularly The total number of occupied states in such a cylinder filled stable. At the lowest conductances these minima occur syn- ␧ 5͒ up to the bulk Fermi energy F is chronously with the conductance jumps, coinciding with the

␧ maxima in the density of states. But when multiple singulari- ͑␧ ͒ϭ ͑␧ ͒ϭ ͵ F ͑␧Ϫ␧ ͒ ␧ ties of the spectrum are very close to each other, forming a N F ͚ Nmn F ͚ LDmn mn d m,n m,n ␧ ͑ Ϸ ͒ mn bunch like those at kFR 8.5, 10.5, 12 etc. the force shows one global extremum. This bunching and the corresponding 2m 1/2 ϭ ͩ e ͪ ͱ␧Ϫ␧ pronounced force extrema mark the fundamental differences ͚ 2L ␲2ប2 mn. m,n between the conductance quantization and shell effects: the former stems from individual spectral singularities, while the The total energy of the electron system is latter is due to the inhomogeneous distribution of the features ␧ in the spectrum ͑caused by geometrical symmetries͒, which ϭ F ␧ ͑␧Ϫ␧ ͒ ␧ E ͚ ͵ L D mn d results in their periodic grouping, or bunching. At this point m,n ␧ mn we cannot say anything about the periodicities of the bunch- 2 2m 1/2 ing as a function of radius. While it is easy to calculate them ϭ ͫ ͩ e ͪ ͑␧ Ϫ␧ ͒3/2ϩ␧ ͑␧ ͒ͬ ͚ L 2 2 F mn mnNmn F . m,n 3 ␲ ប numerically for not too large kFR, an exact analytical ap- proach using the double Poisson summation formula32 is ␮ϭ␧ Assuming F at low temperatures, we can write the ther- complicated, and a much more transparent derivation based modynamic potential on the semiclassical Bohr-Sommerfeld quantization condi- 816 Low Temp. Phys. 27 (99–10), September–October 2001 Yanson et al.

FIG. 14. Density of states ͑DOS͒͑a͒ and force ͑c͒ in the nanowire calculated using Eq. ͑5͒. Smooth backgrounds, subtracted before performing the Fourier transforms, are also shown. Parts ͑b͒ and ͑d͒ show Fourier transforms of the oscillating parts of DOS and the force, respectively. tion is used instead. Even if we plot the density of states ͑or For a circular cross section of the nanowire, the condition for any other quantity dependent on it͒ numerically on a large validity of the semiclassical approximation reads kRӷ1. In- scale, the periodicities are not yet seen with the naked eye stead of finding the discrete quantum mechanical eigenval- ͑Fig. 14a͒. So naturally we would like to take a Fourier trans- ues, one uses the approximate Bohr-Sommerfeld semiclassi- form of exactly calculated thermodynamic quantities ͑5͒ to cal quantization condition: see whether the periodicities are indeed there, and we follow 1 1 the same procedure as we used above for experimental data Ͷ ϭ␯ϩ ␲ប pdx , processing. Thus we simply subtract the smoothed back- 2 2 ground by hand ͑see the smooth curves in Figs. 14a and 14c͒ where ␯ is a positive integer, p is the momentum, and the and perform the Fourier transformation. Both frequency integration is performed over one period of classical motion 6͒ spectra ͑Figs. 14b and 14d͒ give us three prominent peaks of the particle in the plane perpendicular to the wire axis. at 0.63, 0.83, and 0.90. The latter two are very close to each For an axially symmetric circular cylinder, this condition other, and, taking into account the reduced amplitude of the transforms into a one-dimensional quantization condition third peak and its proximity to the second one, we can say along the radius variable, of the form:37 that the oscillating part of the density of states as well as the R force is mainly composed of two major frequencies, 0.62 and ͵ ͱk2Ϫm2/r2drϭ␲͑nϩ␧/4͒, ͑6͒ 0.87. These correspond to the two peaks in the experimental m/k Fourier spectra at 0.64 and 0.89 ͑Fig. 6͒. This is a very im- where kϭͱ2m E/ប,E is the kinetic energy, mϭ...2, portant result, which shows that quantum oscillations in each e Ϫ1,0,1,2,... is the azimuthal quantum number, n is a positive of the quantities of Eq. ͑5͒ possess the principal periodicities integer, ␧/4Ͻ1, and r is the wire radius, which has two turn- of the energy spectrum, and that this seemingly irregular ing points: m/k and R, respectively. Integrating Eq. ͑6͒,we spectrum of electron energies in a cylindrical nanowire can obtain the quantization condition for k: be decomposed into a few basic harmonics. Semiclassical approximation. Much more information kR͑sin ␸Ϫ␸ cos ␸͒ϭ͑nϩ␧/4͒␲, ͑7͒ about the bunching of energy levels, leading to periodicities cos ␸ϭm/͑kR͒. in the spectrum, can be obtained from the semiclassical approach.30,34–36 In this approach we consider an electron as This is an implicit dependence of k on n and m from which a wave packet constructed from many quantum mechanical the semiclassical spectrum can be derived. eigenstates. The wave packet has wave vectors k concen- In the case of a complicated geometry, this approach is trated in the range satisfying the uncertainty principle in the preferred to solving the Schro¨dinger equation exactly, but in space region where we would like to localize our particle. our simple symmetric case we can obtain the spectrum ex- Low Temp. Phys. 27 (99–10), September–October 2001 Yanson et al. 817 actly ͑see above͒ and therefore we shall rather use this rela- conclusion that the peaks in the Fourier spectrum of the os- tion to analyze the behavior of k. Following Balian and cillating part of the force in the nanowire, located at the Bloch,38 we shall look for bunching of the eigenvalues— frequencies of Table I, each correspond to a single semiclas- those intervals of k where the energy levels are very close to sical trajectory. Every trajectory is characterized by two pa- each other. In order to find them, we should consider the rameters ͑t, p͒ and is composed of all eigenstates with quan- variations ␦m and ␦n and find the conditions at which the tum numbers ͑m, n͒ chosen such that ⌬mϭp, ⌬nϭt.In ␦ Ӷ ⌿ corresponding (kR) 1 for some fixed R. As we are look- other words, by grouping together the wave functions mn ing at the semiclassical approximation, we should consider based on this principle ͑⌬mϭp, ⌬nϭt, where p and t are only the part of the spectrum for large k and n where we can fixed͒, one can construct a wave packet or a ‘‘particle’’ trav- ignore ␧. Combining both equations in ͑7͒ and varying m and eling along a classical trajectory of a polygon with p vertices n, we obtain: inscribed in a circle and making t turns around the center. There is an important difference between cluster physics ␦͑ ͒ ␸ϭ␲␦ ϩ␸␦ ␦ ␦ ෈ ͑ ͒ kR sin n m; n, m Z. 8 and that of nanowires. In spherical clusters the spectral ϭ In order to have ␦(kR)ϭ0 for nonzero ␦m and ␦n, the weight of the diametric orbit (t,p) (1,2) is much smaller 7͒ value of ␸ should be a rational fraction of ␲:␸/␲ϭt/p.Of than that of the triangular and square orbits. Then, in order course ␸ has to satisfy Eqs. ͑7͒, which have solutions of the to observe the beating between the latter two, one has to form38 ␸ϭ(t/p)␲ϩO(1/kR), where p and t are two inte- register a large number of oscillations, since the first mini- gers, satisfying the condition pу2t. Since they enter only in mum occurs at about 11–12 maxima due to the closeness of the form of a fraction t/p, we can assume that they are the frequencies. This is not a simple task because of the co-prime. strongly decaying amplitudes, but one can make use of an We can now fix the values of t,pÞ0 and look for the additional feature: the nearly identical intensities of the tri- corresponding periodicity in kR. Substituting ␸ϭt␲/p into angular and square orbits give rise to the missing of one half 39 the same variation ͑8͒, we obtain of the oscillation period at the minimum. In the nanowire, in contrast, because of the less symmetric configuration, the ␲ intensity of the diametric orbit is approximately similar to ⌬͑ ͒ ␸ϭ ͑ ⌬ ϩ ⌬ ͒ kR sin ␳ p n t m . that of the triangular and square orbits ͑see Fig. 14͒. Thus the beating between them is observed at a much higher fre- In order to determine the period of this dependence, we have quency, as can easily be seen in the experimental histograms. ⌬ Þ to find the smallest possible value of (kR) 0. It follows The missing half-period is not observed in the nanowire, from Bezout’s theorem that for integers p and t prime to each because of the small number of oscillations between the ⌬ ⌬ ⌬ ϩ ⌬ other there exist integers n and m such that p n t m nodes in the beating pattern. ϭ ⌬ 1. Thus we obtain the periodicity of the bunches: k The intensities of the spectral contributions of each tra- ϭ ␲ ␸ 2 /(2pR sin ). On the other hand, it is easy to see that jectory are determined by their weight in the semiclassical ϭ ␸ ␸ϭ ␲ L(p,t) 2pR sin , t /p is just the perimeter of a poly- expansion. From Figs. 14b, d one can see that for all pϾ4 gon inscribed into a circle of a radius R having p vertices and and tϾ1 the contributions become negligibly small. This can making t revolutions around the center. Therefore we see that be explained if one recalls that pϭ⌬m, tϭ⌬n, so the our spectrum will contain multiple sets of periodic bunches smaller these are, the more individual energy levels ͑or wave ͑ of levels determined by the choice of p and t or, alterna- eigenfunctions͒ for a given limiting Fermi energy ␧ will be ⌬ ⌬ ͒ F tively, n and m , and each set can be described as a present in the spectrum and contribute to a particular trajec- closed semiclassical trajectory inside a circular cross section tory, increasing its weight in the spectrum. An exact deriva- of a cylinder. Some of them are shown in Table I. This is a tion of this fact can be found in Ref. 38. very general result: it allows one to take a full set of eigen- functions of the system determined by varying two quantum numbers m and n, and using the above-mentioned relations 4.2. Atomic shells to ‘‘create’’ particles in the form of wave packets moving Structure of alkali nanowires: 6-fold symmetry. Re- along certain trajectories ͑polygons inscribed in a circle͒. cently several papers have appeared that consider unconven- Now let us compare the frequencies which we obtain in tional structures for nanowires.40–42 In Ref. 40 a 5-fold the exact quantum mechanical treatment ͑Fig. 14b, d͒ with cluster-like structure was observed for a sodium nanowire the ones determined by semiclassical trajectories. As the through computer simulation. For gold, Kondo and Takay- number of vertices of a polygon with winding number tϭ1 anagi claimed to observed spiral-like shell structures with an inscribed in a circle grows, its perimeter approaches the incremental number of 7 atoms in cross section.42 Molecular circle itself. Thus, the minimum period ⌬(kR)ϭ2␲R/ dynamic simulations for the s-p metals Al and Pb were re- L(p,t)ϭ1 would be for a trajectory which is the circle itself, ported in Ref. 41, where a ‘‘weird’’ spiral-like structure was setting the maximum frequency in the spectrum. The posi- predicted for nanowires with conductances smaller than tions of the first few periods are calculated, and the corre- 10G0 . Some of these weird wires resemble the structures sponding trajectories are shown in Table I. observed in Ref. 42. Note that for thicker nanowires the cal- It is important to note that the values that we obtain for culations predict a close-packed hexagonal structure. the frequencies of the first few orbits coincide perfectly with When searching for an explanation for the structure that the positions of the peaks on the Fourier spectrum of the we have called ‘‘atomic’’ in our experiment, we are again led calculated electronic density of states ͑or force͒ in a perfect by the research on metal clusters. For alkali-metal clusters cylinder ͑Fig. 14b, d͒. We can therefore draw the important produced in vacuum a clear transition has been observed 818 Low Temp. Phys. 27 (99–10), September–October 2001 Yanson et al. between the series of magic numbers given by the closing of ϭ1.34, we find that the beginning of the atomic-shell struc- electronic shells and the other series of magic numbers de- ture in Fig. 9 is described by the shell numbers jϭ4, 5, and termined by the closing of geometric shells of atoms.8–10 The 6, which are closed at g1/2ϭ6.02, 7.36, and 8.70, respec- latter series result from the fact that the clusters have a tively. Apart from the small shift, which may be absorbed in ϭ highly symmetric crystalline shape, and the surface energy of a new choice of the offset j0 0.66, these values correspond a cluster attains a minimum when a new layer of atoms com- to the higher-intensity peaks with indices 7, 13, and 19, and pletely covers the surface. Let us assume that nanowires have five smaller peaks between them. By analogy to the obser- a similar tendency toward crystalline order, and form facets. vation on Ca clusters,10 this suggests an enhanced stability Such faceting has indeed been observed recently in transmis- for nanowires with all six facets covered. Alternatively, one sion electron microscope images of gold nanobridges under can check the self-consistency of the linear relation ͑9͒ by 43 thinning of the bridges by the electron beam current. If assuming that the first fully covered layer with 6-fold sym- many hundreds of wires are drawn, the facets should be ex- metry should contain at least 6 alkali atoms in the cross pressed due to a well-defined minimum in the surface energy. section, taking into account that 1 monovalent atom approxi- Since the periodic pattern extends to large wire diameters, mately corresponds to a single conductance channel and, cor- we first start with the assumption that the lattice structure in respondingly, to a reduced conductance gϷ6. Thus, follow- the wire is that of the bulk metal. The bulk lattice structure of ing formula ͑9͒, for the slope ␣ we obtain ͱ6/(1ϩ0.66) ͓ ͔ potassium is bcc and the lowest energy surfaces are the 110 ϭ1.47, very close to the theoretical value 1.447 and the ex- ͓ ͔ surfaces. A wire with only 110 facets can be formed with perimental slope 1.34. ͓ ͔ ͓ ͔ the wire axis along 100 and 111 . The cross section of the Furthermore, proceeding this way, we obtain the expla- former would be square while that of the latter is a hexagon. nation for the step structure shown in Fig. 13. When we The hexagonal wire clearly has a smaller surface area and successively remove 6 hexagonal facets with the same num- would be preferred. ͑ ͒ ͑ ͒ 1/2 ber on average of atoms see the inset in Fig. 13 , they In order to calculate the slope of g versus the peak should give 6 successive stability peaks, which increase the index i, we approximate the relation between the conduc- conductance g by the same average increment, depending on tance and wire cross section A by the lowest-order term, g the size of the hexagon. While receding the electrodes, the Ϸk2 A/4␲, which is appropriate for k Rӷ1. The distance F F next full atomic shell should give us the next 6 stability between the ͓110͔ atomic layers is hϭa /&, where a is 0 0 facets with an increment decreasing approximately by one the size of the cubic unit cell. We thus obtain atom, i.e., by one conductance quantum, etc. This is just -ͱg 31/4 what is observed in Fig. 13, where the edges of the ‘‘plaץ ϭ k a . teaus’’ coincide with the maximal peaks with indices 7, 13, i ͱ␲ F 0ץ 2 19, and the increment between successive ‘‘plateaus’’ ϭ ␲2 1/3 For a bcc lattice kFa0 (6 ) , and for the slope we obtain amounts to about 1 quantum unit. a value ␣ϭ1.447, which is shown in the inset in Fig. 9 as a One-atomic-layer slab model for stretching nanowire. line labeled ‘‘full atomic shell.’’ This is clearly a much In this paragraph we show that the hexagonal close- higher slope than the one we observe, and even higher than packed structure of a nanowire is compatible with the often- the slope for the electronic shell structure. discussed model of plastic flow of a metallic nanowire dur- Sodium has a low-temperature martensitic phase transi- ing pulling off of the electrodes. tion toward a close-packed atomic structure and it is possible Molecular dynamics simulations performed by Landman that the surface tension favors close packing in nanowires. It et al.1 suggest that plastic flow of a nanometer-sized contact turns out that one can construct a similar shape of a nanow- involves a series of successive structural transformations of a ire, as proposed above for the bcc structure having close small number of atomic layers adjacent to the narrowest part packing. Such a wire would have an fcc lattice, with the wire of the constriction. In each transformation the neck shrinks axis along ͓011͔, and six facets that are perpendicular to as one atomic layer is added. Let us consider for simplicity ͓100͔, ͓11¯ 1͔, ͓¯1¯11͔, ͓¯100͔, ͓¯111¯ ͔, and ͓111¯ ͔. For this the circular symmetric model described in Refs. 20 and 21. arrangement we obtain ␣ϭ35/6␲1/6/213/12ϭ1.427, which is The narrowest part of a wire is responsible for its electrical very close to the value for the bcc structure. conductance. Untiedt et al. propose that a part of the cylin- When we assume that a stable configuration is obtained drical constriction of the junction, which they call the trans- each time when a single facet of the nanowire is completed, formation length ␭, depends on the minimal cross section ␭ϭ␣ n ␣ in analogy to what is observed for some metal clusters ͑e.g., area A as 0A , where 0 is a constant and n is an ͒ for aluminum͒,10,44 then the slope becomes a factor of 6 exponent.8 Depending on the indentation procedure and lower, equal to 0.241, in close agreement with the experi- other parameters of the experiment ͑temperature, elongation ͒ ␣ mental data. Further support for this interpretation comes rate, presence of adsorbates, etc. , 0 and n may have vari- from considering the conductance values for closing of a full ous values. shell. We can write According to Ref. 21, the process of pulling the nanow- ire can be represented schematically as shown in the inset of g1/2ϭ␣͑ jϩ j ͒, ͑9͒ 0 Fig. 15. Starting with a relatively thick bridge of cross sec- ͓ ͑ ͔͒ ␭ where j is the index of a full shell of atoms, with all facets tion A0 see a , the length 0 first transforms elastically, ϭ ͑ ͒ covered, and j0 0.5 is an offset value, which depends some- which is schematically shown in b . The elastic elongation what on the boundary conditions for the electrons. From this corresponds to a plateau in the conductance versus elonga- expression, and using the experimental value ␣ϭ6ϫ0.224 tion curve. Next, it goes through a yielding stage, which Low Temp. Phys. 27 (99–10), September–October 2001 Yanson et al. 819

packed atoms in the nanowire. This is important, because as the nanowire is stretched the crystal facets do not evaporate like in mass spectroscopy but should be built into the thin- ning wire, if we neglect their escape to the electrodes.

5. SUMMARY We have presented evidence for two sets of magic num- bers in metallic nanowires: electronic and atomic, in com- plete analogy with cluster physics. These two sets of oscilla- tions in the conductance histogram compete with each other, as in cluster physics, where the transition from the electronic shell structure to the atomic shell structure is also observed for larger diameters.9 One shell-closing effect is related to the energy of the total volume of electrons, for which the amplitude of the oscillations in the thermodynamic potential decreases as 1/R. The other is due to the surface energy, for which the amplitude of the oscillations is roughly constant as a function of R. The transition between them depends on the parameters of experiment. The atomic-shell oscillations are observed at larger diameters ͑conductances͒ than the elec- tronic ones, but they may overlap substantially. For Li, and FIG. 15. Dependence of the conductance versus elongation for an averaged scan obtained by plotting positions of the atomic shell peaks in the range in many cases for Na, the electronic-shell structure oscilla- 30ϽgϽ250 versus the elongation, which is calculated as iϫh, where i is a tions completely dominate the spectrum. successive integer and h is the interlayer distance. The fit ͑curve͒ of the The periodic peak structure in the histograms is only experimental data points ͑squares͒ by Eq. ͑11͒ gives an exponent nϭ0.51 observed at temperatures well above helium temperature. ϭ ␭ ϭ and a transformation length at g 250 of 0 13.3 nm. Thermal energy is required in order to have sufficient mobil- ity of the atoms allowing the structure to accommodate to the lowest free energy. Potassium has the lowest melting tem- corresponds to a jumplike shrinking of the diameter. Follow- perature among three alkali metals ͑Li, Na, K͒ studied by us. ing Refs. 1, 20, and 28, we assume that the wire relaxes by This means that at a given temperature its atoms have the creating one more atomic layer ͑see Figs. 15b and 15c of the ͒ highest mobility, and for this metal we obtain the largest inset . We take into account the discreteness of the atomic number of oscillations in the conductance histogram. From structure, which consists of close-packed atomic layers with our data we cannot exclude other atomic wire arrangements, thickness h. Then, due to volume conservation, the atoms since they depend strongly on the material studied.41,42 How- from a fraction x of the surface layer with a transformation ␭ ␲ ␭ ever, in the latter studies, beyond a critical radius of the order length and a wire radius R,2 Rx (A)h, should be trans- of three atomic distances the bulk lattice structure is recov- ferred to the new atomic layer along the wire axis ␲R2h ͑see ͒ ered. Therefore, it is likely that the nanowires with diameters Fig. 15 . Equating these two quantities we obtain in the range of the atomic-shell structure oscillations ob- 2␲Rx␭͑A͒hϭ␲R2h,or␭͑A͒ϭR/2x. ͑10͒ served here have a regular atomic stacking structure. On the other hand, the alkali metals Li and Na have low-temperature In this model a proportionality between ␭ and R exists re- martensitic phase transitions toward a close-packed atomic gardless of the presence of the atomic shell effect, as was structure, and it is likely that the surface tension favors close shown in Ref. 21. Taking into account that at elevated tem- packing in the nanowires. peratures the proportionality between piezo voltage and elon- Since we have studied conductance histogram for alka- gation may be violated, we use instead a model to extract a line metals up to high conductances, we can claim that al- scale for the elongation. We assume that the distance be- most all the structure observed originates from favorable tween the conductance peaks of the atomic-shell structure atomic configurations and not from conductance quantiza- corresponds to an elongation hϭͱ8/3r , equal to the dis- 0 tion, unlike the initial few peaks ͑1, 3, 6͒ at low tempera- tance between the close-packed layers of atoms ͑r is the 0 tures, where transmission measurements convincingly show atomic radius͒. In this way we obtain the conductance- degenerate quantized conductance modes.18 However, the fa- versus-elongation curve shown in Fig. 15. Fitting this curve vorable atomic configurations are influenced by the quanti- to the model dependence21 zation of the electronic levels, and only the atomic shell nl 1/n structure is a purely classical effect, but still on the atomic ͑ ͒ϭ ͩ Ϫ ͪ ͑ ͒ g l gmax 1 ␭ , 11 scale. 0 ϭ ϭ ␭ ϭ␭ ϭ where l is the elongation, gmax g(l 0) and 0 (l 0), I. K. Y. is grateful to the Kamerlingh Onnes Laboratory we obtain ␭/RϷ3.1. Here we start the elongation at the con- of Leiden University for its hospitality and for the opportu- ϭ ϭ ductance gmax 250, which corresponds to Rmax 4.26 nm nity to participate in the work presented in this paper. Special ␭ ϭ ␭Ϸ ϭ and 0 13.3 nm. From 3R it follows that x 1/6, which thanks to Prof. L. J. de Jongh for his interest and continuous perfectly correlates with hexagonal symmetry of close- support. 820 Low Temp. Phys. 27 (99–10), September–October 2001 Yanson et al.

*Present address: Dept. of Physics, 510 Clark Hall, Cornell University, Ith- 14 C. J. Muller, J. M. van Ruitenbeek, and L. J. de Jongh, Physica C 191,485 aca, NY 14853 ͑1992͒. 1͒ ϵ 2 ϭ ⍀ Ϫ1 15 One conductance quantum equals G0 2e /h (12907 ) . E. N. Bogachek, A. N. Zagoskin, and I. O. Kulik, Sov. J. Low Temp. Phys. 2͒ ϭ ͑ ͒ From now on we shall use the reduced conductance as g G/G0 , where 16,796 1990 . ϭ 2 16 ͑ ͒ G0 2e /h. J. A. Torres, J. I. Pascual, and J. J. Saenz, Phys. Rev. B 49, 16581 1994 . 17 3͒We shall discuss the transformation length21 in the paragraph concerning B. Ludoph and J. M. van Ruitenbeek, Phys. Rev. Lett. 82, 1530 ͑1999͒. the atomic shell structure. 18 B. Ludoph and J. M. van Ruitenbeek, Phys. Rev. B 61, 2273 ͑2000͒. 4͒The latter corresponds to the frequency 0.866. 19 I. K. Yanson and A. V. Khotkevich, Atlas of Point-Contact Spectra of 5͒The assumption that the Fermi energy of the wire is always equal to that of Electron-Phonon Interaction in Metals, Naukova Dumka, Kiev ͑1986͒. the bulk is, generally speaking, invalid.31 This correction, however, does 20 J. A. Torres and J. J. Saenz, Phys. Rev. Lett. 77, 2245 ͑1996͒. not change the results qualitatively and, since we are only interested in the 21 C. Untiedt, G. Rubio, S. Vieira, and N. Agra¨it, Phys. Rev. B 56, 2154 periodicities of the thermodynamic quantities, it can be ignored for the ͑1997͒. sake of simplicity. 22 O. Yu. Kolesnychenko, O. I. Shklyarevskii, and H. van Kempen, Rev. Sci. ͒ 6 The features at about 1.3, 1.6, and higher are the higher-order harmonics of Instrum. 70, 1442 ͑1999͒. the main frequencies; their presence will be explained later. The minor 23 C. Z. Li, H. Sha, and N. J. Tao, Phys. Rev. B 58, 6775 ͑1998͒. peak at 0.3 is probably a subharmonic due to the size of the interval of 24 Yu. F. Komnik, Fiz. Tverd. Tela ͑Leningrad͒ 10, 312 ͑1968͓͒Sov. Phys. transformation. This ‘‘extra’’ feature is an artifact of our crude method—it Solid State 10, 248 ͑1968͔͒. 35 is absent if one treats the problem exactly. 25 L. J. Lewis, P. Jensen, and J. L. Barrat, Phys. Rev. B 56, 2248 ͑1997͒. 7͒ Any polygon inscribed in a diametric cross section of a sphere has an 26 Tables of Physical Quantities (A Handbook), edited by I. K. Kikoin, At- additional degree of freedom, that of rotation around an axis lying in the omizdat, Moscow ͑1976͒. plane of the polygon. Only the diametric orbit does not have this feature, 27 C. Ho¨ppler and W. Zwerger, Phys. Rev. Lett. 80, 1792 ͑1998͒. and hence its contribution to the spectrum is reduced. 28 A. Stalder and U. Du¨rig, J. Vac. Sci. Technol. B 14, 1259 ͑1996͒. 8͒ ␭ϭ This form is a generalization of the expression const which was 29 S. Blom, H. Olin, J. L. Costa-Kra¨mer, N. Garsia, M. Jonson, P. A. Serena, ϭ adopted in Ref. 28 and corresponds to n 0. and R. I. Shekhter, Phys. Rev. B 57, 8830 ͑1998͒. 30 C. Ho¨ppler and W. Zwerger, Phys. Rev. B 59, R7849 ͑1999͒. 31 J. M. van Ruitenbeek, M. H. Devoret, D. Esteve, and C. Urbina, Phys. ͑ ͒ 1 U. Landman, W. D. Luedtke, N. A. Burnham, and R. J. Colton, Science Rev. B 56, 12566 1997 . 32 ´ ͑ ͒ 248,454͑1990͒. E. N. Bogachek and G. A. Gogadze, Zh. Eksp. Teor. Fiz. 63,1839 1972 ͓ ͑ ͔͒ 2 B. J. van Wees, H. van Houten, C. W. J. Beenakker, J. G. Williamson, Sov. Phys. JETP 36, 973 1973 . 33 L. P. Kouwenhoven, D. van der Marel, and C. T. Foxon, Phys. Rev. Lett. C. A. Stafford, D. Baeriswyl, and J. Bu¨rki, Phys. Rev. Lett. 79,2863 ͑ ͒ 60,848͑1988͒. 1997 . 34 ͑ ͒ 3 D. A. Wharam, T. J. Thornton, R. Newbury, M. Pepper, H. Ahmed, J. E. F. C. Yannouleas and U. Landman, J. Phys. Chem. B 101, 5780 1997 . 35 Frost, D. G. Hasko, D. C. Peacock, D. A. Ritchie, and G. A. C. Jones, C. Yannouleas, E. N. Bogachek, and U. Landman, Phys. Rev. B 57,4872 J. Phys. C 21, L209 ͑1988͒. ͑1998͒. 36 4 J. M. Krans, J. M. van Ruitenbeek, V. V. Fisun, I. K. Yanson, and L. J. de J. Lerme´, Ch. Bordas, M. Pellarin, B. Baguenard, J. L. Vialle, and Jongh, Nature ͑London͒ 375, 767 ͑1995͒. M. Broyer, Phys. Rev. B 48, 12110 ͑1993͒. 5 A. I. Yanson, G. Rubio Bollinger, H. E. van den Brom, N. Agra¨it, and 37 L. D. Landau and E. M. Lifshitz, Course of Theoretical Physics, Vol. 3, J. M. van Ruitenbeek, Nature ͑London͒ 395, 783 ͑1998͒. Pergamon Press ͑1976͒. 6 H. Ohnishi, Y. Kondo, and K. Takayanagi, Nature ͑London͒ 395, 780 38 R. Balian and C. Bloch, Ann. Phys. 69,76͑1972͒. ͑1998͒. 39 J. Pedersen, S. Bjo¨rnholm, J. Borggreen, C. Bre´chignac, K. Hansen, T. P. 7 W. D. Knight, K. Clemenger, W. A. de Heer, W. A. Saunders, M. Y. Chou, Martin, and H. D. Rasmussen, Nature ͑London͒ 353, 733 ͑1991͒. and M. L. Cohen, Phys. Rev. Lett. 52, 2141 ͑1984͒. 40 R. N. Barnett and U. Landman, Nature ͑London͒ 387,788͑1997͒. 8 W. A. de Heer, Rev. Mod. Phys. 65,611͑1993͒. 41 O. Gu¨lseren, F. Ercolessi, and E. Tossati, Phys. Rev. Lett. 80,3775͑1998͒. 9 M. Brack, Rev. Mod. Phys. 65,677͑1993͒. 42 Y. Kondo and K. Takayanagi, Science 289,606͑2000͒. 10 T. P. Martin, Phys. Rep. 273, 199 ͑1996͒. 43 Y. Kondo and K. Takayanagi, Phys. Rev. Lett. 79, 3455 ͑1997͒. 11 A. I. Yanson, I. K. Yanson, and J. M. Ruitenbeek, Nature ͑London͒ 400, 44 T. P. Martin, U. Na¨her, and H. Schaber, Chem. Phys. Lett. 199,470 1440 ͑1999͒. ͑1992͒. 12 A. I. Yanson, I. K. Yanson, and J. M. Ruitenbeek, Phys. Rev. Lett. 84, 5832 ͑2000͒. This article was published in English in the original Russian journal. Repro- 13 A. I. Yanson, I. K. Yanson, and J. M. Ruitenbeek ͑to be published͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Thermoelectric effects in a Luttinger liquid I. V. Krive*

B. Verkin Institute for Low Temperature Physics and Engineering of the National Academy of Science of Ukraine, 47 Lenin Ave., Kharkov 61103, Ukraine I. A. Romanovsky

V. N. Karasin Kharkov National University, 4 Svobody Sq., Kharkov 61077, Ukraine E. N. Bogachek, A. G. Scherbakov, and Uzi Landman

School of Physics, Georgia Institute of Technology, Atlanta, GA 30332-0430, USA ͑Submitted March 19, 2001͒ Fiz. Nizk. Temp. 27, 1110–1122 ͑September–October 2001͒ Thermoelectric effects in a Luttinger liquid ͑LL͒ wire adiabatically connected to the leads of noninteracting electrons are considered. For a multichannel LL a staircase-like behavior of the thermal conductance as a function of chemical potential is found. The thermopower for a LL wire with an impurity is evaluated for two cases: ͑i͒ LL constriction, and ͑ii͒ infinite LL wire. We show that the thermopower is described a Mott-like formula renormalized by an interaction-dependent factor. For an infinite LL the renormalization factor decreases with increase of the interaction. However, for a realistic situation, when a LL wire is connected to the leads of noninteracting electrons ͑LL constriction͒, the repulsive electron-electron interaction enhances the thermopower. A nonlinear Peltier effect in a LL is briefly discussed. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1414571͔

1. INTRODUCTION are neutral, whereas their topological excitations carry charge and spin. Charge and heat transport through a narrow wire whose Since the LL and the FL have qualitatively different ex- width is comparable to the electron Fermi wavelength occur citation spectra, the transport properties of LLs have been the via a finite number of transport channels associated with subject of theoretical interest, and it was shown rather early4 quantization of the electron’s transverse momentum in the that the electric conductance G of an impurity-free infinite ϭ wire. Furthermore, at low temperatures the phase-breaking LL depends on the interelectron interaction, i.e., G gG0 , ␭ ␭ ϭ 2 length, ␳(T), can exceed the length of the wire, ␳(T) where G0 e /h is the quantum of conductance and g is the ϾL, and the electron transport becomes phase coherent. In dimensionless electron–electron interaction parameter of the the Landauer approach1 to such quantum mechanical trans- LL. Subsequent intensive investigations pertaining to trans- port problems the complexity of calculating the relevant port properties of LLs were triggered by the studies of Kane transport coefficients is reduced to a single-particle scattering and Fisher5 and of Glazman et al.,6 who considered the problem, with the transport properties of the electrons de- transport of charge through a local impurity in the LL, find- scribed in terms of the probability for transmission of the ing that for repulsive electron-electron interactions the con- electrons through the effective scattering potential repre- ductance scales with temperature ͑at low temperatures͒ as a sented by the wire. Indeed, this approach, whose implemen- power law G(T)ϳT2/gϪ2; such behavior has been reported tation is often simpler than the use of the Kubo treatment of in recent experiments.7,8 such problems, has proved to be most useful for the descrip- Heat transport in a LL was first considered in Ref. 9, tion of the transport properties of noninteracting electrons where it was shown that in an infinite homogeneous LL the through wires ͑constrictions͒ of reduced dimensions ͑see re- thermal conductance K(T) is not renormalized by the inter- ͒ ϭ ϭ ␲2 2 views in Ref. 2 . actions, i.e., K(T) K0(T) ( /3)kBT/h, while in the pres- It is well known that for strictly one-dimensional ͑1D͒ ence of an impurity K(T)ϳT3. This result, together with the interacting electron systems the Fermi liquid ͑FL͒ descrip- one for the electrical conductance, predicts violation of the tion of the low-energy excitations does not hold. Instead, for Wiedemann-Franz law in a LL. such systems with interactions which leave the electronic The above results, which were derived for effectively spectrum gapless, the corresponding ‘‘long-wavelength’’ infinite LLs, cannot be tested directly in quantum wires con- theory is that of the Luttinger liquid ͑LL͒.3 Unlike the Fermi nected to source and drain leads. To address this issue, the liquid description, where charged excitations are represented transport properties of the LL were considered for a finite 1D by quasiparticles ͑electrons and holes͒, electrons do not wire adiabatically connected to FL leads modeled by 1D res- propagate in an ͑infinite͒ LL. Rather, the excitation spectrum ervoirs of noninteracting electrons. The results obtained for of the LL consists of gapless bosonic excitations ͑charge and such a finite and impurity-free LL wire were found to be spin density waves͒; harmonic oscillations of boson fields qualitatively different from those derived for the infinite LL.

1063-777X/2001/27(9–10)/10/$20.00821 © 2001 American Institute of Physics 822 Low Temp. Phys. 27 (99–10), September–October 2001 Krive et al.

In particular, it was shown that for finite LL wires with adia- lem was used. The finite LL wire is modeled by an effective batic contacts to the reservoirs the electric conductance is not transmission coefficient which in the Landauer-Buttiker ap- renormalized by the interelectron interaction10 and that the proach determines the charge and heat transport between the thermal conductance is significantly suppressed ͑for spinless leads. We predict that the thermopower of a LL with an electrons͒ for a strong repulsive interparticle interaction.11,12 impurity is described by a Mott-like formula—it depends These predictions have a rather simple physical explanation. linearly on the temperature and is proportional to the loga- Since the electrons in the reservoirs are taken as noninteract- rithmic derivative of the bare ͑unrenormalized by the ing particles, one could use the Landauer approach for cal- electron-electron interactions͒ electric conductance. At low Ӷ⌬ Ӎប ͑ culation of the electric and thermal conductances. For an temperatures kBT L s/L L is the length of the LL adiabatic LL constriction the electrons are not backscattered wire, and s is the plasmon velocity͒ the thermopower is not by the confining potential of the wire, and consequently renormalized by the electron-electron interactions, and it is charge is transmitted through the wire with unit probability. described by the well-known formula for the thermopower Therefore, the electric conductance of a LL constriction co- ͑ S0 for a system of noninteracting electrons see, e.g., Ref. incides with the conductance of a single-channel quantum ͒ ӷ⌬ 14 . At temperatures kBT L the interaction renormalizes point contact. In contrast, heat is transported in the LL by the thermopower, and consequently for a strong interaction plasmons ͑charge-entropy separation͒11 which, for strong in- ϳ 2ӷ SLL S0 /g S0 . The renormalization factor is different for teractions, are significantly backscattered at the ‘‘transition’’ spinless and spin-1/2 electrons, and the enhancement of the region between the LL wire and the FL reservoirs, and con- thermopower is more pronounced for spinless particles. sequently heat transport is suppressed. Next, we calculate the thermopower for an infinite LL. The aforementioned studies dealt with spinless electrons Although the situation when the effects of the leads are ex- and a single-channel LL. However, in many real situations cluded appears somewhat artificial from the experimental ͑ ͒ the quantum wires may support several 1D transport chan- point of view, it is useful to elaborate this problem by a nels, and currently thermoelectric effects in LLs remain powerful LL calculation technique. In particular, we note that largely unexplored. In this context, we remark that it has 13 the transport properties of 1D interacting electrons have been been noted that the thermopower of a Hubbard chain, in the studied mostly for an infinite LL, and thus the evaluation of vicinity of a Mott-Hubbard phase transition to a dielectric the thermopower for this case represents an interesting and phase, can be calculated using the Mott formula ͑see, e.g., important theoretical problem. We show that for an infinite Ref. 14͒ for noninteracting fermions. This observation has LL wire with an impurity the thermopower is described by been exploited15 in a derivation of the thermopower of a the Mott formula, S , multiplicatively renormalized by the homogeneous infinite Hubbard chain in the limits when the 0 electron-electron interaction. Hubbard model can be mapped onto a model of spinless For an infinite LL the renormalization factor decreases Dirac fermions. with increase of the interaction, S(gӶ1)ϳgS . This result In light of the above, we report here on studies of heat 0 does not contradict our previous claim, since the two prob- transport through a multichannel LL constriction connected lems under study ͑infinite LL wire and LL wire adiabatically to Fermi liquid leads, as well as investigations of the ther- connected to metallic leads͒ are not identical. In particular, mopower ͑Seebeck͒ and Peltier effect in a LL wire ͑Fig. 1͒. the driving voltage which enters the definition of the ther- First, we study heat transport through a multichannel LL mopower is different for the two cases in question. For an constriction. In this case the thermal conductance as a func- infinite LL it is the voltage drop V across the impurity. In the tion of the chemical potential ␮ demonstrates a staircase-like case of the LL constriction the bias voltage U is defined as behavior. We show that at low temperatures TӶT Ӎបv /L 0 0 the difference of the chemical potentials of the leads, U ͑ is the characteristic velocity, which is determined by the v0 ϭ⌬␮ 16 strength of the confining potential, and L is the length of the /e. It has been shown that for a strong impurity ͑ ͒ ϭ 2 LL wire͒ the steps in the conductance K(␮) are practically weak tunneling V g U. Thus the thermopower of a LL unaffected by electron-electron interactions. On the other wire, when expressed in terms of U, is enhanced by interac- hand, strong interactions suppress the heat conduction at tion. This derivation supports our finding that the strong in- ϳ terelectron interaction strongly enhances the thermopower of temperatures T T0 ; however, the steps are pronounced even in this high-temperature region. Subsequently, we a LL with an impurity. ͑ ͒ evaluate the thermopower for a finite LL wire connected to It is well known see, e.g., Ref. 14 that in the linear- FL leads. In this case a simple physical approach to the prob- response regime the Peltier effect is determined by the same thermoelectric coefficient as the Seebeck effect. However, in the nonlinear regime the Onsager symmetry relations be- tween the transport coefficients cease to be valid, and the у ͑ ͒ Peltier coefficient for eV kBT V is the bias voltage de- scribes an independent thermoelectric phenomenon. We evaluate the nonlinear Peltier coefficient for an impurity- containing LL wire connected to leads. The phenomenologi- cal approach, when the finite LL wire is modeled by an ef- fective transmission coefficient, does not predict the ͑ ͒ FIG. 1. Schematic of a Luttinger liquid LL nanowire of length L, con- renormalization of the nonlinear differential Peltier coeffi- nected to Fermi liquid ͑FL͒ reservoirs that are kept at different temperatures. The impurity ͑scattering potential, denoted by X͒ is placed in the middle of cient by the interaction. the LL wire. The paper is organized as follows. In Sec. 2 the thermal Low Temp. Phys. 27 (99–10), September–October 2001 Krive et al. 823

conductance of a multichannel LL is studied. In Sec. 3 the ␲ប 1/2 v1 v1 v1 thermopower of a LL constriction with an impurity is evalu- g ϵ Ӎͩ ͪ Ӷ1; 1 s U 2ϩ 2 ated in a phenomenological approach. In Sec. 4 bosonization 1 0 v1 v2 technique in conjunction with a tunneling Hamiltonian v method is used for a calculation of the thermopower of an ϵ 2 Ӎͱ р ͑ ͒ g2 v2 /v1 1. 5 infinite LL. In Sec. 5 we investigate the Peltier effect in a s2 Luttinger liquid. The main results are summarized in Sec. 6. Note that for spin-1/2 interacting electrons the Hamil- tonian of a single channel LL is given by Eq. ͑1͒ with N ϭ ϭ ϭ 2 and v1 v2 v. In this case the velocity of the ‘‘spin’’ ϭ 2. INTERACTION-ENHANCED STAIRCASE BEHAVIOR OF mode s2 v is not renormalized by the interaction, i.e., gs THE THERMAL CONDUCTANCE ϭ1. In the following we will see that ‘‘spin’’ channels offer ‘‘easy pathways’’ for heat transport through a LL constric- To calculate the thermal conductance of a multichannel tion. LL wire adiabatically connected to 2D reservoirs of nonin- In the absence of electron backscattering ͑see discussion teracting electrons we will use the multimode LL model de- below͒ the plasmon modes are noninteracting. Consequently, veloped in Ref. 16. The Hamiltonian of the model in the the Landauer approach1 can be used for calculation of the boson representation takes the form thermal conductance. The corresponding expression reads11,12 N p2͑x͒ mn ϭ ͵ ͫ j ϩ j 2͑ Ј͒2ͬ H ͚ dx v j u j N ϱ f ץ ϭ 2mn 2 1 j 1 j ͑ ͒ϭ ͵ ␧␧2ͩ Ϫ B ͪ ͑␧͒ ͑ ͒ ␧ tn , 6ץ K T ͚ d N Th nϭ1 0 U0 ϩ ͵ ͑ ͒ Ј͑ ͒ Ј͑ ͒ ͑ ͒ Ϫ ͚ nin j dxfL x ui x u j x , 1 ϵ͓ ␧ Ϫ ͔ 1 2 i, jϭ1 where f B exp( /kBT) 1 is the Bose-Einstein distribu- ␧ tion function of the plasmons, and tn( ) is the probability of where u j(x) is the displacement operator of the jth mode; plasmon transmission through the nth mode of the LL. As ץ ץЈϵ u j u j / x; p j is the conjugate momentum, with we have said, we assume here that the contacts of the LL to ϭ ប␦ ␦ Ϫ ͓ui(x),p j(y)͔ i ij (x y); n j is the number density of the Fermi liquid reservoirs are adiabatic, which means that ϭ␲ប the electrons in the jth mode and v j n j /m is the corre- there is no backscattering of charged excitations in the LL. ͑ ͒ sponding Fermi velocity, and U0 determines the strength of Formally Eq. 6 represents the thermal conductance of a the electron-electron interaction, which is assumed to be lo- purely bosonic noninteracting system.18 As was shown in Ϫ ϭ ␦ Ϫ ͑ cal: U(x y) U0 (x y). We introduced into the Hamil- Refs. 11 and 12, this formula also applies to an adiabatic no ͑ ͒ ͒ tonian in Eq. 1 a smooth function f L(x) that restricts the electron backscattering LL wire, where the heat is trans- electron-electron interaction to a finite region of length L. ported by bosonic excitations ͑plasmons͒, whose dynamics, The electron reservoirs are modeled as 1D N-channel Fermi in the absence of local scatterers, is described by a quadratic gases and they are represented, in the boson form, by the Hamiltonian. These considerations lead one to conclude that noninteracting part of the Hamiltonian. Eq. ͑6͒ yields the exact thermal conductance of a LL wire in The Hamiltonian in Eq. ͑1͒ is quadratic and can be easily the absence of impurities. However, the plasmons could be diagonalized. In diagonal form it describes N noninteracting backscattered by the ‘‘transition region’’ between the LL and ϭ ‘‘bosonic’’ modes with velocities sn (n 1,...,N) which are the FL reservoirs. Since the width d of the transition regions ␭ Ӷ Ӷ adiabatically transformed into N modes with velocities vn obeys F d L, we can model them as zero-width bound- (nϭ1,...,N). The latter modes correspond to the N noninter- aries located at xϭ0 and xϭL. Consequently, the mismatch acting electron channels in the leads. The plasmon velocities of the plasmon velocities at the boundaries will cause strong 17 sn are determined by the equation backscattering of the plasmons. Thus the transmission coef- ␧ ͑ ͒ ficient tn( ) in Eq. 6 can be obtained by taking the function N v ␲ប f (x) in Eq. ͑1͒ to be of the form f (x)ϭ␪(x)␪(LϪx) ͚ n ϭ ͑ ͒ L L 2Ϫ 2 . 2 ͓where ␪(x) is the Heaviside step function͔ and matching the nϭ1 s vn U0 wave functions of the plasmons at the boundaries. Since ␧ For a two-channel (Nϭ2) case the above equation can there is no channel mixing, tn( ) takes a form analogous to be easily solved, yielding that calculated in Ref. 12: ␧ 1 1 2 ␧ Ϫ1 1 1 ͑␧͒ϭͫ 2ͩ ͪ ϩ ͩ ϩ ͪ 2ͩ ͪͬ ͑ ͒ ϭͱ 2ϩ 2 Ϯ ͱ 2Ϫ 2 2ϩ ␲ប 2 tn cos gn sin , 7 s ͒ ͑u u ͒ ͑u u ͒ ͑2U / ͒ v v , ⌬ ⌬ 1͑2 2 1 2 2 1 2 0 1 2 n 4 gn n ͑3͒ ⌬ ϵប where n sn /L is the characteristic energy scale for the ϭ where finite LL wire, and the plasmon velocities sn (n 1,...,N) are determined by Eq. ͑2͒. Note that for spin-1/2 electrons the

u ͑ ͒ϭv ͑ ͒ͱ1ϩU /͑␲បv ͑ ͒͒. ͑4͒ ‘‘spin’’ mode is not renormalized by interaction, and the cor- 1 2 1 2 0 1 2 (s)ϭ ϭ responding correlation parameters gn 1(n 1,...,N/2); (s)ϭ In the limit of strong interelectron repulsion, i.e., U0 i.e., for the ‘‘spin channels’’ one has tn 1, and the heat ӷ␲ប v1(2) , the interaction parameters of the two-channel transport associated with spin density wave excitations is not ϭ у LL, defined as gn vn /sn , take the form (v1 v2) affected by the electron-electron interaction. 824 Low Temp. Phys. 27 (99–10), September–October 2001 Krive et al.

The expressions given in Eqs. ͑2͒, ͑5͒, and ͑7͒ generalize the problem of heat transport through a single-mode spinless 11,12 LL to a multichannel LL. Now the Fermi velocities vn depend both on the chemical potential ␮ and the ‘‘trans- verse’’ quantum number n which characterizes the quantiza- tion of the transverse electron momentum. For a parabolic ϭ ⍀2 2 confining potential Uc(y) m y /2 the corresponding Ќϭប⍀ Ϫ transverse energy takes the values En (n 1/2) (n ϭ1,...,N), and the Fermi velocity of the nth mode is given by ␮ 1 ␮ 1 1/2 v ϭv ␪ͩ ϩ Ϫnͪͩ ϩ Ϫnͪ , ͑8͒ n 0 ប⍀ 2 ប⍀ 2 ϭͱ ប⍀ where v0 2 /m. The appearance of the step function in the definition of the Fermi velocities of the multimode LL results in a staircase behavior of the electric G(␮) and ther- mal K(␮) conductances as functions of the chemical poten- tial ␮. An important comment concerning Eqs. ͑6͒–͑8͒ is war- ranted here. Note that Eq. ͑7͒ is an exact result for noninter- acting plasmon excitations—that is, when the electrons are not backscattered by the confining potential in the LL con- striction. Such a condition is fulfilled at low temperatures and for chemical potentials satisfying ␮Þប⍀(nϪ1/2). In the vicinity of ␮ϭប⍀(nϪ1/2) an additional electron mode ␲2 2 is converted from an evanescent to a propagating mode. This FIG. 2. The thermal conductance, in units of kBT/3h, plotted as a func- tion of the dimensionless chemical potential ␮/(ប⍀)ϩ1/2 for several val- implies that upon reaching the threshold ␮ for entrance into ␲ប ͑ ͒ ues of the strength of the electron-electron interaction U0 /( vF). In a ˜ ϭ ប ϭ ͑ ͒ ˜ ϭ the contact, the character of the corresponding mode the temperature was taken to be T kBTL/( v0) 0.1, and in b T 10. changes, and in doing so the mode is strongly influenced by the confining potential. Consequently, at such threshold val- ues of the chemical potential the assumption of adiabaticity where G is the electric conductance and L is the cross- of the LL constriction fails, and in calculating the thermal transport coefficient which connects the electric current to conductance the contribution due to electron transport needs the temperature difference for noninteracting particles. These to be considered. However, it is well known that the trans- coefficients can be calculated using a formalism developed ͑ ␭ ͒ in Ref. 19 and adapted in Ref. 18 to the Landauer scheme.1 port of charge through a local of the order of F potential in a LL is strongly suppressed due to plasmon renormaliza- In this approach the transport coefficients are expressed in ␧ tion of the bare scattering potential,9 implying that for suffi- terms of the transmission probability t j( ) for an electron to ciently long wires and for strong electron-electron repulsion arrive at the drain electrode in the jth channel as f ץ the contribution of electron transport to the thermal conduc- N ϱ tance K(T) is small and can be neglected. Therefore, we ͑ ␮͒ϭ ͵ ␧ͩ Ϫ F ͪ ͑␧͒ ͑ ͒ ␧ t j , 10ץ G T, G0 ͚ d conclude that under such circumstances Eq. ͑6͒ is valid for jϭ1 0 practically all values of the chemical potential except at the and very beginning of the steps. We note that at low tempera- N f ␧Ϫ␮ ץ Ӷប k ϱ tures, T v0 /L, the staircase-like behavior of the thermal ͑ ␮͒ϭ B ͵ ␧ͩ Ϫ F ͪ ͑␧͒ ͑ ͒ ␧ t j . 11ץ L T, G0 ͚ d conductance is practically unaffected by electron-electron in- h jϭ1 0 kBT teraction ͑Fig. 2a͒. At high temperatures Tӷបv /L the ther- 0 Here G is the conductance quantum and f (␧Ϫ␮) is the mal conductance, although being suppressed in the case of 0 F Fermi-Dirac distribution function of the electrons in the strong interaction,11,12 still demonstrates a clear staircase be- leads. havior as a function of chemical potential ͑Fig. 2b͒. Equations ͑10͒ and ͑11͒ cannot be applied to an infinite LL, where electrons are not propagating particles and the 3. IMPURITY-INDUCED THERMOPOWER IN A LUTTINGER- conventional scattering problem is ‘‘ill-posed.’’A general ap- LIQUID CONSTRICTION proach for calculating transport coefficients in a system of The thermopower is a measure of the capability of a strongly interacting particles is the Kubo formalism, and a system of charged particles to generate an electromotive recent publication where it was used for calculation of the force when a temperature gradient is applied across the sys- thermopower for a Hubbard chain can be found in Ref. 15. tem. In the linear-response regime it can be represented as a As may be seen from that study, with the Kubo approach it is ratio of transport coefficients, difficult to calculate the thermopower in the whole range of external parameters ͑temperature, interaction strength, den- L͑T,␮͒ sity of particles, etc.͒, and indeed the final analytical expres- S͑T,␮͒ϭϪ , ͑9͒ G͑T,␮͒ sions for the desired quantities were derived13,15 only in the Low Temp. Phys. 27 (99–10), September–October 2001 Krive et al. 825 limits when the Hubbard model can be mapped to a model of the peaks will be universal ͑it will not depend on the con- noninteracting fermions, for which a Mott-type expression crete shape of the confining potential͒. To make more- for the thermopower could be used. definite predictions we need to evaluate the thermopower for To obtain thermopower results pertaining specifically to a quantum wire with a single impurity. the transport properties of systems of strongly interacting Since it is known that in the presence of an impurity the electrons, and to consider quantum-wires thermoelectric ef- conductance of a LL is strongly suppressed, one may naively ␮ will also be stronglyץ/Gץfects which could be tested in experiments, we choose to expect that the thermopower Sϰ invoke at first certain simplified ͑yet physically reliable͒ suppressed in such a wire. However, as we show below, that models of strongly interacting electron systems. Such physi- is not the case. Instead, we find that for strong ͑repulsive͒ cal models of charge transport in LLs of strongly, as well as electron-electron interactions the impurity-induced ther- weakly, interacting electrons were proposed in Refs. 6 and mopower of a LL is significantly enhanced in comparison 20 and were shown to yield the same results as those ob- with the thermopower of a system of noninteracting par- tained from more conventional ͑and rigorous͒ treatments of ticles. LL effects,5,21 through the use of Landauer-like expressions To calculate the thermopower of a finite-length LL in the for estimating the dependence of the conductance on the presence of a local impurity ͑which we place for simplicity at ͒ temperature and on the bias voltage. In this Section and in the middle of the constriction we will model the effective Sec. 5 we use such a phenomenological approach ͑see also transmission coefficient as Ref. 22͒ for studying the Seebeck and Peltier effects in mul- ␣ tichannel LLs. ⌬ eff͑␧͒ϭ ͑␧͒ͩ L ͪ ͉␧Ϫ ͉Ӷ⌬ ͑ ͒ When a LL is connected to FL reservoirs with given t t0 ⌳ for EF L , 12 temperatures and chemical potentials one could make use of ͑ ͒ ͑ ͒ ␧ Eqs. 10 and 11 , with t j( ) now regarded as the probabil- and ity of transmission of the electrons ͑in the jth channel͒ through the effective potential barrier formed by the LL ␧ϪE ␣ piece of the wire. For a wire which is adiabatically connected eff͑␧͒ϭ ͑␧͒ͯ Fͯ ͉␧Ϫ ͉ӷ⌬ ͑ ͒ t t0 ⌳ for EF L . 13 to the leads the transmission coefficient is unity as long as we neglect the backscattering of electrons by the confining ␧ Ӷ potential. For a perfect wire the backscattering effect is ex- Here t0( ) 1 is the bare transmission coefficient deter- ponentially small for practically all values of the chemical mined by the unrenormalized scattering potential ͑we restrict ͒ ⌬ ϭប potential, except at the narrow regions in the vicinity of con- ourselves to a single-mode LL ; L s/L is the character- ductance jumps ͑steps͒ where an additional mode is con- istic low-energy scale ͑s is the plasmon velocity͒, and ⌳ is verted from an evanescent to a strongly propagating mode. the cutoff energy, which for a purely 1D LL is of the order of ␣ Such a physical picture results in a staircase-like behavior of the Fermi energy EF . The exponent depends on the the conductance as a function of the chemical potential and electron-electron interaction strength and is different for 17 is often modeled by abrupt jumps of the electron transmis- spinless and spin-1/2 electrons: sion coefficient from zero ͑reflected mode͒ to one ͑transmit- ted mode͒. This model is too simplified for real quantum 1 U Ϫ1/2 ␣ϭ Ϫ ϭ ϩ 0 ϭ ͑ ͒ point contacts, where the specific features of the confining 2ͩ 1ͪ ; g ͩ 1 ␲ប ͪ for s 0, 14 g vF potential could be important for a correct description of the transition region between the conduction plateaus. However, and for strongly interacting electrons this simple model, which does not depend on the details of the bare scattering poten- 2 2U Ϫ1/2 tial, could be a correct approximation. Indeed, the transmis- ␣ϭ Ϫ ϭ ͩ ϩ 0 ͪ ϭ ͑ ͒ 1; gs 2 1 ␲ប for s 1/2. 15 sion of electrons through a long but finite LL is determined gs vF by an effective scattering potential that includes the effects of electron-electron interactions. This potential for sufficiently The transmission probability teff in Eq. ͑12͒ results in an Ӷ long wires and for temperatures kBT EF quenches all expression for the linear conductance which coincides ͑up to modes whose bare transmission coefficients t0 are not very an irrelevant numerical constant͒ with that obtained in Ref. close to unity ͑see the corresponding discussion in Ref. 22͒. 25 via a renormalization group calculation. In fact, the same Since according to Eqs. ͑9͒–͑11͒ the thermopower S(T,␮) expression has been used6 for estimation of the temperature ␮ we observe that for a multimode LL constriction dependence of the LL conductance in the limit of strongץ/Gץϰ the thermopower vanishes on the conductance plateaus and it interaction (gӶ1); this is also the limit of interest to us, peaks at the conduction steps ͑that is, at the transition regions since for weak interactions LL effects would be much from one conductance plateau to the next͒. The qualitative weaker. distinction of the thermopower in a LL from that evaluated The bare transmission is commonly assumed to be a 23,24 for noninteracting electrons lies in the shape of the ther- smooth function of the energy around EF , i.e., mopower peaks. For strongly interacting electrons a simple tץ -approximation in which the ͑now effective͒ transmission co ͑␧͒Ӎ ͑ ͒ϩ͑␧Ϫ ͒ͩ 0 ͪ ͑ ͒ ␧ . 16ץ efficient is modeled by a Heaviside step function could be a t0 t0 EF EF ␧ϭ quite reliable procedure. Then the temperature behavior of EF 826 Low Temp. Phys. 27 (99–10), September–October 2001 Krive et al.

With this form, Eqs. ͑10͒ and ͑11͒ yield From Eqs. ͑20͒–͑22͒ we observe that the LL effects on the thermopower are most significant for strong interactions, G ͑T͒ϭG t ͑E ͒ ӷ␲ប LL 0 0 F U0 vF , and that they are more pronounced for spinless particles than for spin-1/2 electrons ͑Fig. 3͒. ⌬ ␣ L ͩ ͪ , k TӶ⌬ , Since for the thermopower the interaction dependence ⌳ B L factorizes. Equation ͑20͒ could be readily generalized for the ϫ ͭ ␣ case of wires with dilute impurities, where the average spac- k T ͑ Ϫ 1Ϫ␣͒⌫͑ ϩ␣͒␨͑␣͒ͩ B ͪ ⌬ р Ӷ⌳ 2 1 2 1 ⌳ , L kBT , ing between the impurities is large enough so that the impu- ͑17͒ rities act incoherently. In this case the thermopower will still ͑ ͒ Ͼប be described by Eq. 20 at temperatures kBT sn¯ , where ¯n and is the mean concentration of the impurities. An interesting example is a LL junction made of a perfect LL wire of length ␲2 2 kBT L connected to leads through a potential barrier at the con- L ͑T͒ϭG ͩ ͪ tЈ͑E ͒ LL 0 3e 0 F tacts. The thermopower of such a LL junction for tempera- у⌬ ͑ ͒ ͑ ͒ tures kBT L is described by Eqs. 19 and 20 with the ⌬ ␣ ͑ ͒ 0ϭ 0 0 0ϩ 0 0 L total bare conductance G G1G2/(G1 G2), where G1 ͩ ͪ Ӷ⌬ 0 , kBT L , ͑ ͒ ⌳ and G2 are the bare conductances of the contacts. ϫ ͭ ␣ The thermopower, being the ratio of transport coeffi- 6 k T ͑ Ϫ Ϫ1Ϫ␣͒⌫͑ ϩ␣͒␨͑ ϩ␣͒ͩ B ͪ у⌬ cients, is less affected by interaction than the transport coef- 2 1 2 3 2 , kBT L , ␲ ⌳ ficients themselves ͑Eqs. ͑17͒, ͑18͒͒. It is the prefactors in the ͑ ͒ 18 power-law dependences of G(T) and L(T) on the tempera- ture that determine the dependence of the thermopower on where ⌫(x) and ␨(x) are the gamma function and the Rie- the interaction strength. In the phenomenological approach mann zeta function, respectively. developed above, the quantitative correctness of these coef- From Eqs. ͑9͒, ͑17͒, and ͑18͒ we conclude that at low ficients cannot be proved. Therefore, we conclude that the temperatures k TӶ⌬ the thermopower of a LL constriction B L electron-electron interaction enhances the thermopower of a with an impurity is not renormalized by the interelectron LL wire, and we will attempt to find a more rigorous treat- interactions. Instead it is described by a Mott-type formula ment of the problem. In the next Section we evaluate the for noninteracting electrons,24 thermopower of an infinite LL with an impurity by making -ln G0͑␧͒ use of the bosonization technique when calculating the cur ץ ␲2 k2 ͑ ͒ӍϪ B ͩ ͪ ͑ ͒ -␧ , 19 rent in the wire induced by the bias voltage and by the temץ S0 T 3 e ␧ϭ EF perature difference. where G0(␧) is the corresponding ͑bare͒ conductance of the noninteracting electrons. This finding is not surprising, since 4. THERMOPOWER OF AN INFINITE LUTTINGER LIQUID Ӷ⌬ at kBT L the electrons in the leads determine the transport Let us consider an infinite LL wire with a single impurity properties of the LL constriction. However, at temperatures placed ͑for definiteness͒ at xϭ0 ͑i.e., the middle of the wire; ӷ⌬ kBT L the thermopower, being still a linear function of see Fig. 1͒. It is known that for a LL with repulsive electron- temperature, undergoes a strong multiplicative renormaliza- electron interaction the charge transport through an impurity tion: у⌬ ͒Ӎ ͒ ͒ ͑ ͒ SLL͑T L /kB Cs͑g S0͑T , 20

3 1Ϫ2Ϫ1Ϫ␣ ␨͑␣ϩ2͒ C ͑g͒ϭ ͑␣ϩ1͒͑␣ϩ2͒. s ␲2 1Ϫ21Ϫ␣ ␨͑␣͒

Note that unlike the electric conductance GLL(T) and the cross-coefficient LLL(T), the thermopower SLL(T) does not depend on the cutoff parameter, and therefore the interaction- and spin-dependent factor Cs(g) cannot be absorbed into a definition of ⌳. ϭ ϭ For noninteracting electrons Cs(g 1) 1, and the Mott- type formula ͑Eq. ͑19͒͒ holds ͑as it should͒ for all tempera- Ӷ tures (kBT EF). In the limit of strong interaction U0 ӷ␲ប vF

U ͑ Ӷ ͒ϭ 0 ͑ ͒ C0 g 1 12 ␲3ប , 21 vF FIG. 3. The renormalization parameter C(g) and the dimensionless electron interaction parameter g plotted versus the dimensionless strength of the U0 ͑ Ӷ ͒ϭ ͑ ͒ electron-electron interaction U /(␲បv ) for spinless ͑solid line͒ and spin- C1/2 g 1 6 ␲3ប . 22 0 F vF 1/2 ͑dashed line͒ electrons. Low Temp. Phys. 27 (99–10), September–October 2001 Krive et al. 827 is sharply suppressed at low temperatures. Therefore, the LL can expect the contribution (JT) to the charge current in- is ‘‘split’’ by the impurity into two disconnected semi-infinite duced by the temperature gradient. The Hamiltonian given segments, and the charge current through the impurity can be by Eq. ͑25͒ with a constant bare tunneling amplitude does evaluated with the use of the tunneling Hamiltonian method. not allow one to evaluate this contribution. To obtain the We start with the Hamiltonian temperature-induced current we have to take into account the finite size of the barrier. We can do it by modifying the ϭ ϩ H ͚ H0,m Ht , tunneling Hamiltonian. The modified Hamiltonian includes mϭ1,2 extra terms containing the derivatives of the field operators: where H describes two (mϭ1,2) identical semi-infinite 0,m ϩ H ϭ ͚ ͑␭ ⌿ ͑0͒⌿ ͑0͒ϩh.c.͒ parts of the LL wire. In the bosonic form it reads t 0 2,r2 1,r1 r1 ,r2 sប 2 Ϫ1 2 ⌽ ͒ ͔. ͑23͒ ϩ ץ͑ ϩg ͒ ⌰ ץH ϭ ͵ dx͓g͑ ⌿ ͑0͒ ץm 8␲ x m x m ϩ ͚ ͕Ϫiប␭ ͓r ⌿ ͑0͒,0 1 1 2,r2 x 1,r1 r1 ,r2 ϭ Here s is the plasmon velocity, g vF /s is the LL correlation ϩ ⌿ ͑0͒⌿ ͑0͔͒ϩh.c.͖. ͑27͒ ץ ⌽ ⌰ Ϫr parameter, m(x) is the displacement field, and m(x) is the 2 x 2,r2 1,r1 field complementary to ⌽ (x), obeying the commutation re- m Here ͉␭ ͉ is a small additional parameter (͉␭ ͉p ϳ͉␭ ͉). lation ͑see, e.g., Ref. 25͒ ͓⌰ (x),⌽ (xЈ)͔ 1 1 F 0 m m Notice that this form of the Hamiltonian corresponds to a ϭ2␲i␦ sgn(xϪxЈ). The tunneling Hamiltonian is mmЈ tunneling amplitude which depends upon the momentum of 0 ϩϱ the tunneling electron ͗p ͉T͉p ͘ϭ␭ ϩ␭ r q ϩ␭ r q , ϭ ͵ ͵ ͓͗ ͉ ˆ ͉ ͘␺ϩ͑ ͒␺ ͑ ͒ 2 1 0 1 1 1 1 2 2 Ht dx1 dx2 x2 T x1 2 x2 1 x1 where q ϭp Ϫr p is the momentum of the electron to- Ϫϱ 0 m m m F ward the Fermi level. ϩ͗ ͉ ˆ ͉ ͘␺ϩ͑ ͒␺ ͑ ͔͒ ͑ ͒ Now the total electron current through the barrier can be x1 T x2 1 x1 2 x2 , 24 written in the form ␺ ␺ϩ ͑ ͒ where m( ) is the electron annihilation creation opera- m ϩϱ tor, the index m labels two identical semi-infinite segments ϭ ͉␭ ͉2 ͵ ͑ ͒͗⌿ ͑ ͒⌿ϩ ͘ J 2ie 0 ͚ dt sin eVt 2,r t 2,r ͉ ˆ ͉ Ϫϱ 2 2 of the LL wire, and ͗x2 T x1͘ is the tunneling matrix element r1 ,r2 in the coordinate representation, i.e., the amplitude for the ϫ͗⌿ϩ ͑ ͒⌿ ͘ϩ ប͑␭ ␭ 1,r t 1,r 2ie 0 1* process of electron tunneling from the point x1 to the point 1 1 x2 . ϩϱ ϩ␭ ␭ ͒ ͵ ͑ ͒ Let us introduce the ‘‘slow’’ annihilation and creation 0* 1 ͚ dt cos eVt Ϫϱ operators of two types—for right- and left-moving electrons: r1 ,r2 Ϫ ␺ (x)ϭeipFx⌿ (x)ϩe ipFx⌿ (x). At first we suggest m m,R m,L ϩ ϩ ⌿ ͑t,x͒⌿ ͒͗͘⌿ ͑t͒⌿ ͯ͗͘ ץ that our contact is pointlike. Then one can simplify the tun- ϫ͑r 1 x 1,r1 1,r1 2,r2 2,r2 neling Hamiltonian and write it in the form x→0 ϩϱ ϭ ␭ ⌿ϩ ͒⌿ ͒ϩ ͒ ͑ ͒ Ϫ ប͑␭ ␭*ϩ␭*␭ ͒ ͵ ͑ ͒ Ht ͚ ͑ 0 ͑0 1,r ͑0 h.c. , 25 2ie 0 1 0 1 ͚ dt cos eVt 2,r2 1 r r Ϫϱ r1 ,r2 1Ј 2 ⌿ ⌿ϩ ͑ where m,r ( ) is the operator of annihilation cre- ϩ ϩ ⌿ ͑t,x͒⌿ ͒͗͘⌿ ͑t͒⌿ ͯ͘ , ͑28͒͗ ץ m m,rm ϫ͑r 2 x 2,r 1,r ͒ ͑ 2 2,r2 1,r1 1 ation of an electron from the mth half of the wire for right- → ͑ ͒ ϭϩ ͑ ͒ x 0 moving R electrons rm 1, for left-moving L electrons ϭϪ ͒ where ͗...͘ denotes the thermal average, and ⌿ (t) are the rm 1 . m,rm ␭ field operators in the Heisenberg representation, ⌿ We assume that the bare tunneling amplitude 0 is small. m,rm Then the tunneling rate of electrons through the barrier can ϵ⌿ (0). The correlation functions in Eq. ͑28͒ can be m,rm be obtained to leading order from Fermi’s ‘‘golden rule.’’ calculated by making use of the bosonization formula The total rate of electrons from the left ͑‘‘1’’͒ LL to the right ͑‘‘2’’͒ LL is of the form ͑see, e.g., Ref. 27͒ 1 ϩ Ϫ ͓ ⌽ ͑ ͒ϩ⌰ ͑ ͔͒ ⌿ ͒ϭ i rm m x,t m x,t /2 ͑ ͒ m,r ͑x,t U e . 29 m ͱ ␲ m,rm 2␲ 2 a ⌫ ϭ ͉͗ ͉ ͉ Ј Ј͉͘2 12 ប ͚ E1E2 Ht E1E2 ϳប ϩ Ј Ј Here a is the cutoff parameter (a vF /EF), and Um,r is E1ЈE2ЈE1E2 m the unitary raising operator, which increases the number of eV electrons on the branch r by one particle but does not affect ϫP ␦ͩ EЈϩEЈϪE ϪE Ϫ ͪ , ͑26͒ m 12 1 2 1 2 2 the bosonic excitations. We will not specify its form, since this operator enters the formulas we are studying only in the where P is the probability of finding the system in the state ϩ 12 combination UU ϭ1. Now the bosonic fields ⌽ (x,t) and ͉E E ͘, and V is the bias voltage. The standard evaluation m 1 2 ⌰ (x,t) are in the Heisenberg representation. ͑see below͒ of the tunnel current J(V,T)ϭe͓⌫ (V,T) m 12 In our case we have to impose a boundary condition on Ϫ⌫ (V,T)͔ results in the well-known expression for the 21 the displacement field ⌽ (x)atxϭ0 to account for the conductance G(T) of an LL with an impurity.5 m semi-infiniteness of each segment of the LL wire, i.e., Let us assume now that the temperatures of the left (T1) ⌽ ͒ϭ⌽ ͒ϭ ͑ ͒ and right (T2) parts of the wire are different. In this case one 1͑0 2͑0 0. 30 828 Low Temp. Phys. 27 (99–10), September–October 2001 Krive et al.

⌰ ͑ ͒ ͑ ͒ 2 2 2/gϪ2 Besides this, the boson fields m(x) in Eqs. 23 , 29 satisfy ͉␭ ͉ e 1 ␲k Ta ͑ ͒Ӎ 0 ͩ Ϫ1 ͪͩ B ͪ the boundary condition G T ␲ប3 2 B ,g ប , 4 vF 2 vF 1 ϭ ⌬ ͑ ͒ ͑ ͒ 0͒ϭ0. ͑31͒ JT kB TL T , 37͑ ⌰ ץ j͑xϭ0͒ϭ 2␲ x m ␲ e 3 Ϫ The boson fields obeying the boundary conditions Eqs. ͑30͒, L͑T͒Ӎ ͑␭ ␭*ϩ␭*␭ ͒k TBͩ ,g 1 ͪ 2ប3v3 0 1 0 1 B 2 ͑31͒ in the momentum representation take the form F ␲k Ta 2/gϪ2 ϩϱ 2s 1/2 o ϫͩ B ͪ ⌰ ͑ ͒ϭ ͵ ͩ ͪ ͑ Ϫ ϩ͒ ͩ p ͪ ប . m x i dp o b p b p cos x , vF Ϫϱ g p s Here B(x,y)ϭ⌫(x)⌫(y)/⌫(xϩy) is the beta function. ϩϱ 2sg 1/2 o Equation ͑36͒ coincides with the one found in Ref. 28. It ⌽ ͑ ͒ϭ ͵ ͩ ͪ ͑ ϩ ϩ͒ ͩ p ͪ ͑ ͒ m x dp o b p b p sin x , 32 predicts the power-law dependence of conductance on tem- Ϫϱ p s perature. Equation ͑37͒ is a new result. From Eqs. ͑36͒ and ϩ where b p and b p are the standard bosonic annihilation and ͑37͒ one easily gets the thermopower ϩ ϭ␦ 0ϭ ͉ ͉ creation operators (͓b p ,b ͔ p,pЈ); s p is the energy 2 Ϫ pЈ p k ␲2 B͑3/2,g 1͒ T 2 of bosonic excitation with momentum p. ͑ ͒ϭϪ B ͑␭ ␭*ϩ␭*␭ ͒ S g ͑ Ϫ1͒ ͉␭ ͉2 0 1 0 1 . With the help of Eqs. ͑29͒ and ͑32͒ it is straightforward e B 1/2,g 0 vF ͑ ͒ to evaluate the correlation functions. In the vicinity of the 38 contact (xϳ0) one gets the desired correlator For noninteracting electrons (gϭ1) Eq. ͑38͒ has to trans- form into the Mott formula, Eq. ͑19͒. This allows us to relate ϩ 1 ␭ ␭ ͗⌿ ͑x,t͒⌿ ͘Ӎ the parameters 0 , 1 of the tunneling Hamiltonian to the m,r m,rm ␲ m 2 a conductance and its derivative at the Fermi energy ͑ ϩ ͒ ␲ ␹ 1/2 1/g rm 0 Gץ Tm 2 1 1 ϫͫ ͬ ͑␭ ␭*ϩ␭*␭ ͒ϭ ͯ ͑ ͒ ␧ , 39ץ 1ϩiv ␹/a sinh͑␲T ␹͒ ͉␭ ͉2 0 1 0 1 0 F m vF 0 G ␧ϭ EF ͑ Ϫ ͒ 1 ␲T ␩ 1/2 1/g rm ϫͫ m ͬ where G0 is the bare ͑unrenormalized by interaction͒ con- ϩ ␩ ͑␲ ␩͒ , 1 ivF /a sinh Tm ductance. Thus, the thermopower of an infinite LL takes the ͑33͒ form ln G0͑␧͒ 3g ץ ␹ϭ Ϫ ␩ϭ ϩ ͑ ͒ ␲2g k2 T where t x/s and t x/s. By substituting Eq. 33 ͑ ͒ϭϪ B ͯ ϭ ͑ ͒ ␧ ϩ S0 . 40ץ S g ϩ ͒ ͑ into Eq. 28 we find the total electron current. In the linear- 2 g e ␧ϭ 2 g → Ϫ ϭ⌬ → EF response approximation V 0, T1 T2 T 0, the voltage-induced (JV) and temperature-induced (JT) currents We showed that the electron-electron interaction in 1D take the form systems modeled by a Luttinger liquid multiplicatively

2/g renormalizes the thermopower S0 of the Fermi liquid. For an ͉␭ ͉2 2 ϱ ˜ 0 e t Tt infinite Luttinger liquid the renormalization factor decreases J ϭ8i V ͵ dt ͫ ͬ , V ͑ ␲ ͒2 Ϫϱ ͑ ϩ ͒2/g with increasing interaction. At first glance this result, Eq. 2 a 1 ivFt/a sinh͑˜Tt͒ ͑40͒, contradicts the conclusion derived in the previous Sec- ͑34͒ tion. Notice, however, that the two problems in question are 16i␲e k ⌬T ϱ dt not equivalent. It is well known, for instance, that the depen- ϭ ͑␭ ␭ ϩ␭ ␭ ͒ B ͵ JT 0 1* 0* 1 dence of the conductance on the interaction strength is dif- ͑ ␲ ͒2 Ϫϱ 2 a vF ˜Tt ferent for an infinite LL and for a finite LL wire connected to ͑ ͒ 2/g reservoirs of noninteracting electrons see, e.g., Ref. 10 .To Ϫ2/g ˜ vFt Tt relate the two problems under study we will follow the con- ϫͩ 1ϩi ͪ ͫ ͬ ͓˜Tt cosh͑˜Tt͒Ϫ1͔ siderations presented in Ref. 16. In that paper it was shown a sinh͑˜Tt͒ that for a LL wire adiabatically connected to electron reser- Ϫ1 voirs the voltage drop V across the strong impurity ͑no elec- vFt ϫͫ˜Tt cosh͑˜Tt͒Ϫͩ 1ϩi ͪ ͬ. ͑35͒ tron tunneling͒ is connected to the voltage drop U measured a on the leads by the simple relation Vϭg2U. This formula is ˜ ϵ␲ ប ϭ ϩ the manifestation of the Coulomb blockade phenomenon. Here T kBT/ , where T (T1 T2)/2 is the mean tem- perature. Physically it is evident that in the limit of strong interaction 2ϳប 2Ӷ The integrals in Eqs. ͑34͒ and ͑35͒ look very compli- g vF /e 1 the shift of the chemical potentials in the ⌬␮ ϭ cated. Fortunately we are interested only in the limit a→0. leads ( L eU) cannot change significantly the charge In this case the asymptotics of the above integrals can be densities in the LL wire ‘‘split’’ into two parts by a strong impurity potential. So, to relate ͑at least qualitatively͒ the easily found. Both currents JV,T are power-law functions of ϵ␲ ប Ӷ thermopower S(g) evaluated for infinite LL to the ther- the small dimensionless parameter Y kBTa/ vF 1. The leading terms in the asymptotics Y→0 are mopower SLL(g) of a LL smoothly connected to the leads of noninteractive electrons we first of all have to replace the ϭ ͒ ͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ 2 JV VG͑T , 36 voltage V in Eqs. 26 , 28 , 34 , and 36 by g U. Then Low Temp. Phys. 27 (99–10), September–October 2001 Krive et al. 829

Ӷ ϳ 2 g 1 for the current as a function of voltage at low temperatures SLL(g) S(g)/g ——→ S0 /g. This means that for a real situation, when the voltage drop U is measured between the (T→0). Since the backscattering of the electrons by a local leads, the interaction enhances the thermopower. This sup- impurity in an infinite LL leads to a power-law dependence 5 ports our claim based on the calculations done in the phe- of the electric current on the voltage, it may be expected, nomenological approach. Notice that there is still a discrep- and is indeed found in our model, that for a finite LL this ancy ͑by a factor gϪ1 ӷ1͒ between the above estimates and behavior would cross over to ordinary Ohmic JϪV behavior Ӷ⌬ Eq. ͑40͒ in the limit of strong interaction. This inconsistency for eV L . However, the analysis given in Ref. 30 re- could be attributed to the qualitative nature of our estima- vealed the occurrence of additional oscillations of the current tions based on the phenomenological model ͑Sec. 3͒. as a function of the bias voltage, which do not appear in our model. Underlying these oscillations is the multiple scatter- ing of the plasmon by the impurity potential and at the 5. NONLINEAR PELTIER EFFECT IN A LUTTINGER LIQUID boundaries of the LL, and the phase of these oscillations is According to the Thompson relation for the cross- sensitive to the position of the impurity. While our approxi- coefficients of the 2ϫ2 matrix of transport coefficients in the mation scheme does not reveal these mesoscopic oscilla- linear response theory, the Peltier coefficient ⌸(T,V) ͑de- tions, one may expect that such fine structure in the JϪV fined as the ratio of heat current to the electric current in the characteristics would be obliterated upon averaging over the absence of a temperature gradient across the system͒. position of the impurity. With the above assumptions, and using Eq. ͑12͒ in Eq. JQ ͑ ͒ ⌸͑T,V͒ϭͩ ͪ , ͑41͒ 42 , we obtain for the differential electric conductance at J Ӷ e ⌬Tϭ0 kBT eV, ␣ ⌬ ץ ⌸ϭϪ obeys the relation kBTS, where S is the thermopower. Je L ϭG t ͑E ͒ͩ ͪ for eVӶ⌬ , ͑44͒ V 0 0 F ⌳ Lץ It is rather easy to verify that this relation also holds for a LL Ӷ if eV kBT, and thus the linear Peltier coefficient in the LL can be described using Eqs. ͑17͒–͑22͒. In the nonlinear re- and ␣ ⌬ ץ gime, eVӶk T, the Onsager symmetry relations between B Je L ϭG t ͑E ͒ͩ ͪ for eVу⌬ . ͑45͒ V 0 0 F 2⌳ Lץ -the transport coefficients cease to be valid. For noninteract ing electrons the nonlinear Peltier effect has been studied in Ref. 29, and here we remark on its behavior for a LL with an In a similar fashion we obtain for the heat current at Ӷ impurity. kBT eV ␣ ⌬ J e eV 2ץ In the Landauer-Buttiker approach the electric and heat Q Ӎ Ј͑ ͒ͩ ͪ ͩ L ͪ Ӷ⌬ ͑ ͒ t0 EF ⌳ for eV L , 46 ץ currents between reservoirs of noninteracting electrons at ␮ V h 2 fixed temperatures and chemical potentials 1(2) are given by18,19 and ␣ J e eV 2 eVץ G ϱ ͑ ͒ϭ 0 ͵ ␧ eff͑␧͓͒ ͑␧͒Ϫ ͑␧͔͒ ͑ ͒ Q Ӎ Ј͑ ͒ͩ ͪ ͩ ͪ у⌬ ͑ ͒ Je T,V d t f 1 f 2 . 42 t0 EF for eV L . 47 ⌳V h 2 2ץ e 0 ͑ ͒ ͑ ͒ 1 ϱ From Eqs. 44 – 47 it is readily seen that within the J ͑T,V͒ϭ ͵ d␧ teff͑␧͒͑␧Ϫ␮͓͒ f ͑␧͒Ϫ f ͑␧͔͒, ͑43͒ framework of our calculations the nonlinear Peltier coeffi- Q h 1 2 0 cient for a symmetric LL constriction with an impurity where placed at the middle of the LL wire does not depend on the ␧Ϫ␮ Ϫ1 interelectron interactions, and the differential Peltier coeffi- 1͑2͒ ͑ Ӷ ͒ ͑␧͒ϭͫ ͩ ͪ ϩ ͬ cient is given by at kBT eV f ͒ exp 1 1͑2 k T ␧͒ ץ 2 ץ ץ B JQ / V 1 eV ln t0͑ are the distribution functions of the electrons in the reser- ⌸͑V͒ϵ Ӎ ͩ ͪ ͩ ͪ . ͑48͒ ␧ץ V e 2ץ/ Jץ ␮ ϭ␮Ϯ e ␧ϭE voirs, 1(2) eV/2 for a symmetric LL wire, and V is the F voltage drop across the wire. In the following we will use We remark, however, that an influence of the interelec- Eq. ͑12͒͑as in Sec. 3͒ to model the transmission probability eff tron interactions on the Peltier coefficient may occur for t (E) for a finite LL with an impurity placed in the middle asymmetric LL wires or when the aforementioned mesos- of the wire. copic oscillations are included. Prior to proceeding with our analysis we note that the JϪV characteristics of a finite LL connected to FL reservoirs 6. CONCLUSIONS were studied in Refs. 30 and 22 using different approaches. In Ref. 22 the current–voltage dependence was calculated In this paper we have used physically motivated models using a qualitative physical approach, similar to that em- to investigate the heat transport through a multichannel LL ployed by us in the present study, while a more rigorous wire and also the thermopower and Peltier effect in a single- treatment of charge transport through a finite LL with an channel LL with an impurity. impurity, based on renormalization group analysis, was ͑i͒ For a multichannel LL wire, we predict that electron- elaborated in Ref. 30. Unlike the linear-response transport electron interactions would stabilize the staircase-like behav- regime, where the above two approaches arrived at similar ior of the thermal conductance K(T,␮) as a function of the results, in the nonlinear regime they yield different behaviors chemical potential ͑which can be controlled through the use 830 Low Temp. Phys. 27 (99–10), September–October 2001 Krive et al. of a gate voltage͒. For strongly interacting particles the 3 F. D. M. Haldane, J. Phys. C 14, 2585 ͑1981͒; Phys. Rev. Lett. 48, 1840 ␮ϭ␮ ͑1981͒. jumps in the thermal conductance at each value n at 4 which a new propagating channel is allowed to enter the W. Apel and T. M. Rice, Phys. Rev. B 26, 7063 ͑1982͒. 5 ͑ ͒ constriction remain sharp even at comparatively ‘‘high’’ tem- C. L. Kane and M. P. A. Fisher, Phys. Rev. Lett. 68,1220 1992 . 6 L. I. Glazman, I. M. Ruzin, and B. Z. Shklovskii, Phys. Rev. B 45, 8454 peratures. ͑1992͒. ͑ii͒ For a perfect ͑impurity-free͒ LL wire the ther- 7 F. P. Milliken, C. P. Umbach, and R. A. Webb, Solid State Commun. 97, mopower ͑Seebeck coefficient͒ vanishes on the conductance 309 ͑1996͒; A. M. Chang, L. N. Pfeiffer, and K. W. West, Phys. Rev. Lett. plateaus and it peaks sharply at the conductance jumps. We 77, 2538 ͑1996͒. 8 also considered the thermopower effect for a single-channel M. Bockrath, D. H. Cobden, J. Lu, A. G. Rinzler, R. E. Smalley, ͑ ͒ ͑ ͒ LL constriction with an impurity placed at the middle of the L. Balents, and P. L. McEuen, Nature London 397, 598 1999 . 9 C. L. Kane and M. P. A. Fisher, Phys. Rev. Lett. 76,3192͑1996͒. constriction. For this system the Mott expression for the ther- 10 D. L. Maslov and M. Stone, Phys. Rev. B 52, R5539 ͑1995͒;V.V.Pono- Ӷ⌬ ϭប mopower holds at low temperatures kBT L s/L, where marenko, Phys. Rev. B 52, R8666 ͑1995͒; I. Safi and H. J. Schulz, Phys. s is the plasmon velocity and L is the length of the LL wire. Rev. B 52, R17040 ͑1995͒. Ͼ⌬ 11 R. Fazio, F. W. J. Hekking, and D. E. Khmelnitskii, Phys. Rev. Lett. 80, However, at kBT L the thermopower is multiplicatively renormalized by the electron-electron interactions. The effect 5611 ͑1998͒. 12 ͑ ͓͒ of this renormalization is predicted to be more pronounced I. V. Krive, Fiz. Nizk. Temp. 24,498 1998 Low Temp. Phys. 24,377 ͑1998͔͒; Phys. Rev. B 59, 12338 ͑1999͒. for spinless particles than for spin-1/2 electrons. This conclu- 13 H. J. Schulz, Int. J. Mod. Phys. B 5,57͑1991͒. sion is supported by an evaluation of the thermopower for an 14 J. M. Ziman, Principles of the Theory of Solids, Cambridge University infinite LL with an impurity by the tunnel Hamiltonian Press, Cambridge ͑1986͒. method. The Peltier coefficient ⌸(T,V) of a LL wire, in the 15 C. A. Stafford, Phys. Rev. B 48, 8430 ͑1993͒. 16 R. Egger and H. Grabert, Phys. Rev. B 58, 10761 ͑1998͒. linear-response regime, is determined by the thermopower, 17 ͑ ͒ ⌸ ϭϪ ͑ K. A. Matveev and L. I. Glazman, Physica B 189, 266 1993 . (T) kBTSLL(T). Unlike the thermopower Seebeck co- 18 U. Sivan and Y. Imry, Phys. Rev. B 33,551͑1986͒. ͒ efficient the nonlinear Peltier coefficient is found in our 19 E. N. Bogachek, I. O. Kulik, A. G. Shkorbatov, Fiz. Nizk. Temp. 11, 1189 model to be unaffected by the interelectron interactions, and ͑1985͓͒Sov. J. Low Temp. Phys. 11, 656 ͑1985͔͒. thus it is determined by the energy dependence of the bare 20 D. Yue, L. I. Glazman, and K. A. Matveev, Phys. Rev. B 49, 1966 ͑1994͒. 21 probability of transmission through the wire. A. Furusaki and N. Nagaosa, Phys. Rev. B 47, 4631 ͑1993͒. 22 M. Jonson, P. Sandstrom, R. I. Shekhter, and I. V. Krive, Superlattices We gratefully acknowledge support of this work from Microstruct. 23, 957 ͑1998͒. 23 the National Research Council through the Twinning Pro- H. van Houten, L. W. Molenkamp, C. W. J. Beenakker, and C. T. Foxon, Semicond. Sci. Technol. 7, B215 ͑1992͒. gram 1999–2000. I. V. K. thanks the School of Physics at the 24 E. N. Bogachek, A. G. Scherbakov, and U. Landman, Phys. Rev. B 54, Georgia Institute of Technology for its hospitality. The re- R11094 ͑1996͒. search of E. N. B., A. G. S., and U. L. was also supported by 25 A. Furusaki and N. Nagaosa, Phys. Rev. B 54, R5239 ͑1996͒. the U.S. Department of Energy, Grant No. FG05- 26 A. O. Gogolin, A. A. Nersesyan, and A. M. Tsvelik, Bosonization and 86ER45234. Strongly Correlated Systems, Cambridge University Press ͑1998͒. 27 G. L. Ingold and Yu. V. Nazarov, in Single Charge Tunneling, edited by H. Grabest and M. H. Devoret, NATO ASI series B, V. 294, Plenum Press, * E-mail: [email protected] p. 21. 28 C. L. Kane and M. P. A. Fisher, Phys. Rev. B 46, 15233 ͑1992͒. 29 E. N. Bogachek, A. G. Scherbakov, and U. Landman, Solid State Com- 1 R. Landauer, IBM J. Res. Dev. 1, 223 ͑1957͒; Philos. Mag. 21, 863 mun. 108, 851 ͑1998͒; Phys. Rev. B 60, 11678 ͑1999͒. ͑1970͒. 30 V. V. Ponomarenko and N. Nagaosa, Phys. Rev. B 56, R12756 ͑1997͒. 2 M. Buttiker, Semicond. Semimet. 1, 223 ͑1957͒; 35,191͑1992͒;Y.Imry, Introduction to Mesoscopic Physics, Oxford University Press, New York This article was published in English in the original Russian journal. Repro- ͑1997͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Magnetic susceptibility of La0.875Sr0.125MnO3 under hydrostatic pressure to 1 GPa A. E. Petrova,* E. S. Itskevich, V. A. Ventcel’, V. F. Kraidenov, and A. V. Rudnev

L. V. Vereshchagin Institute of High Pressure Physics, Russian Academy of Sciences, 142190 Troitsk, Moscow District, Russia ͑Submitted March 6, 2001͒ Fiz. Nizk. Temp. 27, 1123–1127 ͑September–October 2001͒

The magnetic susceptibility of single-crystal La0.875Sr0.125MnO3 is measured under hydrostatic pressure up to 1 GPa. The Curie temperature TC and the temperatures TM and TOO of quantum phase transitions are determined. The pressure dependence of these transition ϭ ϭ ϭ temperatures is determined: dTC /dP 18 K/GPa, dTM /dP 16 K/GPa, and dTOO /dP 22 K/GPa. The phase diagram obtained demonstrates that an applied pressure can have the same effect on the phase transitions as a change in the degree of doping x.©2001 American Institute of Physics. ͓DOI: 10.1063/1.1414572͔

A great interest in perovskite manganates has arisen in with the signal recorded at the fundamental modulation fre- connection with the discovery of colossal magnetoresistance quency, equal to 19 Hz.19 The amplitude of the modulating in these materials. The ‘‘double exchange’’ theory proposed signal was 800 A/m. The pressure was created in a fixed- in Refs. 1 and 2 was later supplemented to include the pressure chamber.20 The pressure-transmitting medium used electron–phonon interaction and the resulting Jahn–Teller was a 1:1 kerosene–oil mixture. The modulating coil was deformations of the lattice.3 These disruptions of the lattice placed on the housing of the chamber, and the measuring coil are often accompanied by structural transitions and the for- was placed inside the chamber with the sample. As the cham- mation of small polarons. Manganates have complex phase ber was cooled, the pressure was measured by a manganese diagrams involving structural, orbital, spin, and charge resistance manometer, and the temperature by a copper– orderings.4,5 Several orderings can occur simultaneously, and ͑copperϩiron͒ thermocouple placed inside the high-pressure they can have a static or a dynamic character. Experimentally chamber. The rate of cooling did not exceed 1 K/min. obtained phase diagrams in the temperature and external As the temperature was lowered, the pressure in the pressure remain extremely rare for manganates.6 chamber decreased.21 The critical points on the P –T phase The effect of pressure on the thermopower and resistiv- diagram are shown with allowance for the drop in pressure. ϭ ␹ ity of a La1ϪxSrxMnO3 sample with x 0.125 was investi- In Fig. 1, which illustrates the temperature dependence (T), gated previously.7 There have been many studies of samples the pressure corresponds to room temperature. The pressure with this special composition.5,7–16 For an ordered distribu- dependence of the emf of the thermocouple is small. For tion of the Sr impurity over the volume of the crystal, a example, at 1 GPa for a ͑AuϩFe͒–Cu thermocouple at doubling of the lattice period should occur, and that could liquid-helium temperatures the change is less than 2% ͑Ref. lead to the formation of a crystalline superstructure and, as a 22͒. consequence, to the appearance of interesting new effects. The La0.875Sr0.125MnO3 were furnished by the laboratory The P –T phase diagram has its greatest complexity near that of A. M. Balbashov. The quality of the samples was estab- concentration.17 lished by a detailed x-ray structural analysis.23 Ͼ ϭ At higher concentrations (x xC 0.18) the samples are found in a ferromagnetic metallic state in which no magnetic RESULTS transitions below the Curie temperature T have yet been C Figure 1 shows the results of the measurements of the observed. Indeed, a study18 of the temperature dependence of temperature dependence of the magnetic susceptibility ␹ for the magnetic susceptibility ␹ on samples of La Ϫ Sr MnO 1 x x 3 two single-crystal samples at different pressures. The mea- with xϭ0.2 and 0.3 under hydrostatic pressure up to 1.8 GPa surements were made under identical conditions except for did not reveal any features in the behavior of ␹(T). How- the difference in pressure. The change of the amplitude of the ever, a clear understanding of the nature of magnetic transi- signal may be due to mechanical stresses or to annealing tions upon changing temperature has not yet been achieved. during the pressure cycling. For constructing the phase dia- We have therefore measured the temperature dependence of gram the important thing is the position of the characteristic the magnetic susceptibility under hydrostatic pressures of up points and not the amplitude of the signal. On cooling below to 1 GPa on the same La Sr MnO samples that were 0.875 0.125 3 200 K a rapid growth of the susceptibility begins, and then used for the thermopower and resistivity measurements in one observes two anomalous peaks, after which the suscep- Ref. 7. tibility ␹(T) follows the usual trend for a ferromagnet. As in Ref. 15, we attribute the sharp rise in ␹(T) to a transition MEASUREMENT TECHNIQUES from the paramagnetic to the ferromagnetic state. The Curie The magnetic susceptibility of single-crystal samples of temperature TC was determined from the inflection point on ␹ ϭ La0.875Sr0.125MnO3 was measured by a modulation method the (T) curve. At standard pressure one has TC 187 K,

1063-777X/2001/27(9–10)/4/$20.00831 © 2001 American Institute of Physics 832 Low Temp. Phys. 27 (99–10), September–October 2001 Petrova et al.

␹ ͑ ͒ FIG. 1. Magnetic susceptibility versus temperature T for two single-crystal samples of La0.875Sr0.125MnO3 a and b of arbitrary orientation at different pressures. The pressure values are given for room temperature. which agrees with the data of Refs. 5,7,14, and 16. With mum of ␣(T) is observed increases at practically the same ϭ ␹ increasing pressure TC increases at a rate of dTC /dP 18 rate as the temperature TM of the maximum of (T): 16 and K/GPa. 17 K/GPa, respectively. This suggests that the features on the The maximum on the ␹(T) curve at TϷ150 K is appar- ␹(T) and ␣(T) curves may be due to the same transition. ϭ ently due to a second-order phase transition from a conduct- The nature of the transition at TM 175–180 K is not ing ferromagnetic phase to a ferromagnetic insulator.6,15 It is completely clear. It may be due to phase separation into a customary in the literature to denote this temperature as conducting and a nonconducting phase. The conducting ͑ ͒ TOO . With increasing pressure TOO increases at a rate of phase is usually identified with ferromagnetic FM polarons, ϭ ͑ ͒ dTOO /dP 22 K/GPa. which are nucleated already in the paramagnetic PM phase Ͼ ͑ ͒ The origin of the maximum at 180 K at standard pres- at T TC Ref. 24 and whose existence in the PM and an- sure will be discussed in detail below. We denote the tem- tiferromagnetic ͑AFM͒ matrices has been detected over a 25–27 perature at which this maximum is observed by TM . With wide temperature range, 4.2–300 K. Apparently the ϭ increasing pressure dTM /dP 16 K/GPa. growth of the thermopower below TM as the resistivity con- 7 The pressure dependence of the temperatures TC , TOO , tinues to decrease is also indicative of phase separation of ͑ ͒ and TM are presented on the P –T diagram see Fig. 2 . the sample. Optical measurements in Ref. 16 revealed a bi- furcation of the peak of the phonon mode at a frequency of Ϫ1 DISCUSSION OF THE RESULTS 350 cm , due to the bending of the MnO6 octahedra, start- ing at a temperature of Tϭ180 K. This bifurcation increased The presence of features on the ␹(T) curve ͑see Fig. 1͒ as the temperature was lowered. The authors of Ref. 16 near 150 and 180 K suggests that two quantum magnetic attributed this effect to phase separation. transitions occur in addition to the transition from the para- Ferromagnetic ordering makes the Jahn–Teller distor- ϭ ͒ magnetic to the ferromagnetic state (TC 187 K . Similar tions of the lattice energetically unfavorable. When the tem- ␹ ϭ ϭ behavior of (T), with peaks at TOO 149 K and TM 175 perature is lowered below the Curie point, the Jahn–Teller K at atmospheric pressure, was observed on a distortions begin to decrease strongly.25,28 The decrease con- La0.875Sr0.125MnO3 sample in Ref. 15. tinues below the temperature TM , until a structural transition ␹ ϭ 7,14,15 ͑ Figure 3 shows the temperature dependence of and of OЈ–O* occurs at TOO 150 K in Ref. 5 a value TOO 7 the thermopower ␣, which has a minimum at TϷ175 K. ϭ160 K was found͒. The O* phase is very close to the With increasing pressure the temperature at which the mini- high-temperature orthorhombic phase O but with somewhat larger Jahn–Teller lattice distortions, the value of which re-

FIG. 2. Temperatures of the magnetic transitions versus applied pressure. The solid lines show our measurements (xϭ0.125), and the dotted lines the FIG. 3. Temperature dependence of the susceptibility ␹ and thermopower ␣ data of Ref. 6 (xϭ0.14). ͑Ref. 7͒ at standard pressure. Low Temp. Phys. 27 (99–10), September–October 2001 Petrova et al. 833 mains practically unchanged as the temperature is lowered spins and to their localization, i.e., to a ferromagnetic insu- further. This structural transition is accompanied by a strong lator phase. The dependence of the bending angle of the rearrangement of the electronic structure: the ferromagnetic Mn–O–Mn line on the degree of doping x in the compounds ͑ ͒ metal FMM is transformed into a ferromagnetic insulator La1ϪxSrxMnO3 in the orthorhombic structure was studied in ͑FMI͒. Ref. 9. In addition to the peak on the temperature dependence of In the presence of Jahn–Teller distortions the influence the magnetic susceptibility, which we also observed in Refs. of pressure on TC is enhanced significantly because of the 9 and 15, the presence of a magnetic transition at TOO optical vibrations of the oxygen ions.30 This leads to a broad- ϭ150 K is also evidenced by anomalies in the behavior of ening of the effective band Weff and to an increase in TC . the magnetic moment M, which were observed in Refs. 5,12, ӷ The system undergoes a transition from strong (J Weff)to and 14. The start of the localization of the carriers at T weak (JӶW ) coupling, where J is the exchange energy. ϭ150 K is confirmed by the temperature dependence of the eff Increasing the Sr concentration in La Ϫ Sr MnO increases resistivity, which increases rapidly as the temperature is low- 1 x x 3 the average ionic radius, an effect analogous to an external ered below 150 K.7,10,12,14 pressure, i.e., an increase in the covalency of the bonds, On the basis of neutron measurements made in Ref. 8, it W , and T . was concluded that the metal–insulator transition at T eff C ϭ ϭ A calculation using our data and previously published TOO in samples with x 0.125 is due to the formation of a commensurate polaron-ordered phase. In such a phase, data shows that for matching of the characteristic values of ϩ ϭ Mn–O layers containing only Mn3 ions and having Jahn– TOO , TM , and TC for samples with x 0.125, 0.14, and ⌬ ϭϩ Teller distortions like those in the OЈ phase alternate with 0.15, a change x 0.01 corresponds to a pressure in- undistorted layers in which polarons ͑Mn4ϩ ions with holes crease ⌬Pϳ0.7 GPa. localized on them͒ form a regular charge-ordered square lat- Our results on the pressure dependence of TC agree and tice. Resonance x-ray scattering on such samples at 150 K with the data of Ref. 7 to within the error limits (dTC /dP has revealed the appearance of a static antiferromagnetic or- ϭ18 and 16 K/GPa, respectively͒. As the degree of doping x 11 bital ordering of hybridized eg orbitals. As the temperature increases, the pressure derivative initially remains unchanged is lowered, the AFM type of orbital ordering gives way to a ͑16 K/GPa for xϭ0.15) and then drops sharply in the FM Ͼ ͑ ϭ 18 FM type, while the spin ordering remains of the FM type. phase for x xC 2 K/GPa for x 0.4–0.5). Our results on The AFM type of orbital ordering lowers the electron kinetic the temperatures TOO and TC are very close to the data of energy Ekin and greatly weakens the double exchange. The Ref. 6, obtained on a sample with xϭ0.14. Both of these ferromagnetism of the spins is then due entirely to superex- temperatures increase substantially with increasing applied 12 change, which is favored by the orbital ordering. The ab- pressure, showing that pressure, like doping ͑for xр0.15), sence of double exchange promotes localization of the carri- tends to stabilize the ferromagnetic polarons. ers and a FMM–FMI transition. The question of the influence of charge ordering on the dielectric properties of doped manganates was discussed in Ref. 13. A calculation by the Hartree–Fock model gave two stable solutions for a La0.875Sr0.125MnO3 sample, with pos- CONCLUSIONS sible orbital and charge orderings. Even in this case, how- ever, the FM state is realized only in the presence of a finite 1. The temperature dependence of the magnetic suscep- р ͑ ͒ lattice deformation, which persists at T TOO Ref. 28 .Itis tibility of La Ϫ Sr MnO single crystals with xϭ0.125 has ͑ 1 x x 3 this deformation that can give rise to a superstructure e.g., a peaks at two temperatures. The first peak, at TϭT ϭ150 ͒ OO doubling of the lattice period and to the opening of a gap in K, is attributed by the authors to a FMM–FMI transition. the carrier spectrum. ϭ ϭ The second peak, at T TM 180 K, is attributed to mag- In La1ϪxSrxMnO3 compounds, because of the different netic ordering in the FM matrix. equilibrium lengths of the La–O and Mn–O bonds, the cubic 2. Under hydrostatic pressure to 1 GPa the temperatures lattice is distorted and is transformed to rhombohedral or of the magnetic transitions T , T , and T increase at orthorhombic. The Mn–O–Mn line is bent at an angle ␸. OO M C rates of 22, 16, and 18 K/GPa, respectively. This means that Hydrostatic pressure decreases the Mn–O distance, the lat- pressure stabilizes the ferromagnetic phase. tice asymmetry, and the angle ␸. This leads to an increase in the exchange integral of neighboring Mn ions and in the 3. The P –T phase diagrams obtained here and in Ref. 6 hopping matrix element. Pressure also decreases the angle demonstrate that a change in the degree of doping x and an applied pressure P can have an identical effect on the quan- between the t2g spins of the neighboring Mn ions, so that in the temperature region near T the double exchange and the tum phase transitions. C ´ number of ferromagnetic clusters increase, and that is ac- The authors thank E. L. Nagaev and L. I. Koroleva for a companied by a decrease in resistivity. The growth of the helpful discussion of the results, and M. I. Baneeva for ferromagnetic phase is accompanied by a decrease in the analysis of the phase composition of the samples. Jahn–Teller lattice distortions.29 With decreasing tempera- This study was supported by Grant No. 00-02-16019 ture this leads to an earlier structural transition OЈ–O* and from the Russian Foundation for Basic Research, as part of to the appearance of orbital ordering. Because of the spin– the Russian Federation Program on High-Temperature Su- orbit interaction, orbital ordering leads to FM ordering of the perconductivity. 834 Low Temp. Phys. 27 (99–10), September–October 2001 Petrova et al.

*E-mail: [email protected] 15 V. Skumryev, J. Nogues, J. S. Mu´˜noz, B. Martı´nez, R. Senis, J. Fortcu- berta, L. Pinsard, A. Revcolevschi, and Y. M. Mukovskii, Phys. Rev. B 62, ͑ ͒ 1 C. Zener, Phys. Rev. 82,403͑1951͒. 3879 2000 . 16 2 P.-G. de Gennes, Phys. Rev. 118, 141 ͑1960͒. J. H. Jung, K. H. Kim, H. J. Lee, J. S. Ahn, N. J. Hur, T. W. Noh, M. S. 3 A. J. Millis, P. B. Littlewood, and B. J. Shraiman, Phys. Rev. Lett. 74, Kim, and J.-G. Park, Phys. Rev. B 59, 3793 ͑1999͒. 5144 ͑1995͒; A. J. Millis, B. J. Shraiman, and R. Mueller, Phys. Rev. Lett. 17 J.-S. Zhou, J. B. Goodenough, A. Asamitsu, and Y. Tokura, Phys. Rev. 77,175͑1996͒. Lett. 79, 3234 ͑1997͒. 4 A. Urushibara, Y. Moritomo, T. Arima, A. Asamitsu, G. Kido, and 18 Y. Moritomo, A. Asamitsu, and Y. Tokura, Phys. Rev. B 51, 16491 ͑1995͒. Y. Tokura, Phys. Rev. B 51, 14103 ͑1995͒. 19 A. G. Budarin, V. A. Ventcel’, O. A. Voronov, and A. V. Rudnev, 5 H. Kawano, R. Kajimoto, M. Kubota, and H. Yoshizawa, Phys. Rev. B 53, Izmeritel’naya Tekhnika, No. 4, 66 ͑1982͒. 2202 ͑1996͒; 53, R14709 ͑1996͒. 20 E. S. Itskevich, Prib. Tekh. E´ kper., No. 4, 148 ͑1963͒. 6 J.-S. Zhou and J. B. Goodenough, Phys. Rev. B 62,3834͑2000͒. 21 V. A. Ventcel’, E. S. Itskevich, A. E. Petrova, and A. V. Rudnev, Fiz. 7 ͑ ͒ E. S. Itskevich and V. F. KraŽdenov, Fiz. Tverd. Tela St. Petersburg 43, Tverd. Tela ͑St. Petersburg͒ 37, 351 ͑1995͓͒Phys. Solid State 37,191 1220 ͑2001͓͒Phys. Solid State 43,1267͑2001͔͒. ͑ ͔͒ 8 1995 . Y. Yamada, O. Hino, S. Nohdo, R. Kanao, T. Inami, and S. Katano, Phys. 22 E. S. Itskevich and V. F. Kra denov, Prib. Tekh. E´ kper., No. 6, 164 ͑1978͒. ͑ ͒ Ž Rev. Lett. 77,904 1996 . 23 A. M. Balbashov and S. K. Egorov, J. Cryst. Growth 52,498͑1981͒. 9 D. N. Argyriou, J. F. Mitchell, C. D. Potter, D. G. Hinks, J. D. Jorgensen, 24 J. M. De Teresa, M. R. Ibarra, P. A. Algarabel, C. Ritter, C. Marquina, and S. D. Bader, Phys. Rev. Lett. 76, 3826 ͑1996͒; J. F. Mitchell, D. N. J. Blassco, J. Garcı´a, A. del Moral, and Z. Arnold, Nature ͑London͒ 386, Argyriou, C. D. Potter, D. G. Hinks, J. D. Jorgensen, and S. D. Bader, ͑ ͒ ͑ ͒ 256 1997 . Phys. Rev. B 54, 6172 1996 . 25 10 L. Pinsard, J. Rodrı´guez-Carvajal, A. H. Moudden, A. Anane, A. Rev- D. Louca, T. Egami, E. L. Brosha, H. Ruder, and A. R. Bishop, Phys. Rev. ͑ ͒ colevschi, and C. Dupas, Physica B 234–236, 856 ͑1997͒. B 56, R8475 1997 . 26 11 Y. Endoh, K. Hirota, S. Ishihara, S. Okamoto, Y. Murakami, A. Nishizawa, M. Hennion, F. Moussa, G. Biotteau, J. Rodrı´guez-Carvajal, L. Pinsard, T. Fukuda, H. Kimura, H. Nojiri, K. Kaneko, and S. Maekawa, Phys. Rev. and A. Revcolevschi, Phys. Rev. Lett. 81, 1957 ͑1998͒. 27 Lett. 82, 4328 ͑1999͒. G. Allodi, R. De Renzi, and G. Guidi, Phys. Rev. B 57, 1024 ͑1998͒. 12 H. Nojiri, K. Kaneko, M. Motokawa, K. Hirota, Y. Endoh, and K. Taka- 28 E. L. Nagaev, Phys. Status Solidi B 186,9͑1994͒. hashi, Phys. Rev. B 60, 4142 ͑1999͒. 29 X. Xiong, B. Dabrowsky, O. Chmaisen, Z. Bukowski, S. Kolesnik, 13 T. Mizokawa, D. J. Khomskii, and G. A. Sawatzky, Phys. Rev. B 61, R. Dybzinski, and C. W. Kimball, Phys. Rev. B 60, 10186 ͑1999͒. R3776 ͑2000͒. 30 V. Laukhin, J. Foutcuberta, J. I. Garcı´a-Mu´˜noz, and X. Obradors, Phys. 14 S. Uhlenbruck, R. Teipen, R. Klingeler, B. Buchner, O. Friedt, M. Hucker, Rev. B 56, R10009 ͑1997͒. H. Kierspel, T. Niemoller, L. Pinsard, A. Revcolevschi, and R. Gross, Phys. Rev. Lett. 82,185͑1999͒. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Relaxation processes in amorphous films of monoatomic metals V. M. Kuz’menko* and B. G. Lazarev†

Kharkov Physicotechnical Institute National Research Center, ul. Akademicheskaya 1, 61108 Kharkov, Ukraine ͑Submitted July 23, 2001͒ Fiz. Nizk. Temp. 27, 1128–1137 ͑September–October 2001͒ The relaxation of the electrical resistance in amorphous films of Bi, Yb, Be, and Mn is investigated at low temperatures. It is shown that this process during an isothermal hold is ordinarily a simple exponential process with a single characteristic relaxation time. In amorphous metal–hydrogen films a subbarrier tunneling of hydrogen atoms apparently occurs. It is conjectured that this is the cause of the delay in the onset of superconductivity at Tϭ4.2 K after completion of the condensation of amorphous Be–H films. © 2001 American Institute of Physics. ͓DOI: ͓DOI: 10.1063/1.1414573͔

1. INTRODUCTION for Mn͒ investigated in this study ranges from 13 to 70 K. The relaxation behavior of the resistance of amorphous films Amorphous metallic alloys, obtained by the rapid was studied in the temperature interval 4.2– Tcr . Naturally, at quenching of a melt or deposition of a vapor on a cold sub- such low temperatures the relaxational changes in the resis- strate, are configurationally frozen far from internal equilib- tance are small. However, first, the high accuracy of the mea- rium. If they have sufficient atomic mobility, they can lower surements makes it possible to reveal the main behavioral their free energy through some change in the parameters of regularities of this change, and, second, as will be shown the amorphous structure. Such structural relaxation can sub- below, the tiny relaxational changes of the resistance are stantially alter many properties. The main types of experi- sometimes accompanied by large changes in the supercon- ments on the study of relaxation processes in amorphous ducting and kinetic parameters of the films. materials deal with relaxation of stress, volume, enthalpy, viscosity, electrical properties, and superconducting 2. METHODOLOGICAL NOTES properties.1–5 The links between the various changes in prop- erties remain little studied. There have been a few studies in It is known that the only way of obtaining amorphous which the kinetics of different processes ͑e.g., relaxation of films of pure monoatomic metals is condensation of their the enthalpy and volume2 or relaxation of the electrical re- vapors in an ultrahigh vacuum on a substrate cooled by liq- sistance and elastic constants6͒ have been investigated on the uid helium. The details of the technique used in this experi- same material and in the same temperature interval, and ment have been described by us previously.12,14 The tech- these have shown a significant similarity in the relaxation nique ensures a residual pressure of Ϸ10Ϫ10 Pa in the film- behavior of the quantities measured, attesting to the exis- condensation vessel prior to the start of condensation. tence of a unified mechanism for their change. In metallic Although the pressure was not varied during condensation, ͑ amorphous alloys, in addition to the irreversible relaxation the parameters of the films obtained the resistance, Tcr , usually due to a change in topological structure, a reversible critical thickness,12 superconducting transition temperature ͒ relaxation, due to a change in local chemical order, has also Tc , etc. demonstrate that our films are every bit as pure as been observed.3 those obtained in a controlled vacuum with a residual gas The atomic basis of structural relaxation is most easily pressure of 1.33ϫ10Ϫ8 – 1.33ϫ10Ϫ9 Pa.15 understood in amorphous monoatomic metals, in which there In addition to studying the processes of resistance relax- is no influence from atoms of another species. In particular, it ation in pure amorphous films of the metals listed above, we has been shown calorimetrically that reversible relaxation is also investigated the influence of hydrogen ͑up to 50 at.%͒ not observed in glasses consisting of only a metal or a on these processes in Be, Bi, and Yb films. The hydrogen metalloid.3 was introduced into the amorphous films of these metals by Systematic studies of relaxation processes in amorphous condensing them at a partial pressure of pure hydrogen films of monoatomic metals at temperatures close to liquid Ϸ4.67ϫ10Ϫ5 Pa ͑and partial pressures of the components of helium temperature have been extremely rare. Nevertheless, the residual air of less than 10Ϫ10 Pa͒. The technique of ob- a few experimental papers on this topic are known.7–11 taining the metal–hydrogen films is described in detail in The present paper is devoted to a preliminary study of Ref. 14. the relaxational change in the resistance of amorphous films The resistance of the samples obtained was measured of Bi, Ga, Be, V, Yb, and Mn both during their heating to using a potentiometer in a four-probe scheme. The tempera- temperatures slightly below the crystallization temperature ture of the samples was determined by a platinum resistance Tcr and during an isothermal hold at lower temperatures. The thermometer. The accuracy of the temperature measurements value of Tcr depends on the purity and thickness of the amor- in the interval 4.2–20.4 K was 0.05–0.1 K; at higher tem- phous layer of metal12,13 and for the amorphous films ͑except peratures it was 0.5 K or better. The average temperature of

1063-777X/2001/27(9–10)/8/$20.00835 © 2001 American Institute of Physics 836 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Kuz’menko and B. G. Lazarev

TABLE I. Percent relaxational decrease of the resistance ͑␦͒ of amorphous films condensed on a liquid-helium cooled substrate, as a result of their Ͻ heating to Th Tcr .

␦ Metal d,nm Th ,K Tcr ,K ,% Bi 10 20 21 0.2 V 16 34 40 0.55 Yb 6 12.5 19 0.39 Yb 26 9 13 0.09 Be–H 30 55 60 0.79 Be–H 71 58 65 0.29 FIG. 1. The temperature dependence of the electric resistance of a bismuth film, including the crystallization process ͑a͒; the relaxational change in the resistance of this same film in the amorphous state ͑b͒͑see explanation in text͒. from the film during condensation is inversely proportional to its thickness.17 The heat removal from the surface of thick films is slower, and so there is a greater possibility for relax- the Yb, Bi, Ga, Be, Mn, and V films at the time of their ational processes to take place during condensation. For the condensation was measured in auxiliary experiments with metals investigated in the present study the values of ␦ are the aid of a film thermometer deposited directly on the sub- small, usually only a few tenths of a percent ͑see Table I͒ as strate prior to condensation of the metal to be investigated; it a consequence of the small difference between the tempera- did not exceed 5, 5.5, 6, 10, 15, and 18 K, respectively.16 The ture of the film during its condensation and the crystalliza- superconductivity of the pure amorphous films of Bi, Ga, and tion temperature, which sets a ceiling on the temperature at Be during their condensation confirms that their average which relaxation processes can be studied in the homoge- temperature did not exceed their respective values of Tc . neous amorphous state. Amorphous Mn films condensed at Ͻ Ϸ 18 The heating of the samples to the crystallization temperature T 15 K are an exception; for them Tcr 380 K. was done at a rate of 0.5–1 K/min. To fix the relaxation Figure 2a shows the relaxational decrease of the resis- processes more completely, the approach to the temperature tance of an amorphous Mn film Ϸ100 nm thick at T of the isothermal hold was made at a much higher rate. Film ϭ4.2 K as a function of the heating temperature. For the thicknesses greater than 60 nm were determined from the maximally relaxed film the value of ␦ is Ϸ10%. The relax- optical density of the films, and the thickness of the thinner ational decrease of the resistance on heating of amorphous films was measured by an interferometric method. Mn films is accompanied by an appreciable change in the Hall constant RH . This is shown in Fig. 2b, where the values ϭ 3. EXPERIMENTAL RESULTS of RH of the film at T 4.2 are plotted as a function of the heating temperature. The existence of relaxation processes in amorphous Relaxation processes in amorphous films of pure metals films of pure metals is clearly manifested even during their are clearly manifested even during an isothermal hold at T heating, starting at the very lowest temperatures. Figure 1a i shows the temperature dependence of the resistance R on heating of an amorphous film of bismuth Ϸ10 nm thick. The sharp increase of the resistance at a temperature Ϸ21 K ϭ ( Tcr) is due to crystallization of the sample. The part of the Ͻ curve at T Tcr characterizes the amorphous state of the film. Figure 1b shows on an enlarged scale the irreversible change of the resistance of this sample on heating to Tϭ19 K ͑curve 1͒ and the reversible path of R(T) in the interval 11–19 K ͑curve 2͒. Curve 2 characterizes the temperature dependence of the resistance of the relaxed film. Curve 1 reflects both the negative temperature coefficient of resistance of this particu- lar sample and the lowering of R with increasing temperature owing to structural relaxation. The relaxational decrease of the resistance of an amor- phous film can be characterized by the parameter ␦ ϭ⌬ R/R0 , where R0 is the resistance of an as-condensed film at a temperature above the superconducting transition ͑for Bi and Be͒ or at Tϭ4.2 K ͑for V, Yb, and Mn͒;1͒ ⌬R ϭ Ϫ R0 RT , where RT is the resistance of the film at the same temperature after it has been heated to a certain temperature Ͻ Ͻ Th Tcr or after an isothermal hold at Ti Tcr . The relaxational decrease of the resistance is manifested particularly clearly during heating of the thinnest films FIG. 2. Relaxational change in the resistance ͑a͒ and Hall constant ͑b͒ of an (Ͻ5 nm) ͑see Ref. 9, for example͒. This behavior may be amorphous manganese film at Tϭ4.2 K as functions of the heating tempera- due, in particular, to the fact that the rate of heat removal ture Th . Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Kuz’menko and B. G. Lazarev 837

teristic relaxation time ͑for a given hold temperature͒, and n is an exponent (р1) that depends on the temperature of the isothermal hold.2͒ The relaxational change in the resistance ϭ of amorphous films of vanadium at Ti 4.2 K, shown in Fig. 3a, is satisfactorily described by Eq. ͑1͒ with nϭ1 ͑Fig. 3b͒. A deviation from Eq. ͑1͒ is observed only in the initial stage of the relaxation process, t→0, and in the stage where the system is approaching equilibrium. Ytterbium films behave in a qualitatively similar way during a hold at 4.2 K. How- ever, the parameter ␦ characterizing the relaxational decrease in the resistance is much smaller for Yb than for V. The value of ␦ varies from 0.13% for a Yb film of thickness Ϸ3nmto 0.002–0.004% for films thicker than 15 nm. It is clear that in Yb films the relaxation processes are able to take place at the time of condensation. However, even such a small relax- ational change of the resistance can lead to a catastrophic effect in ytterbium films of subcritical thickness—to the ϭ ͑ ͒ spontaneous onset, at Ti 4.2 K, of avalanche explosive crystallization of the entire sample. We attribute this to the formation, in a localized region of the film, of a ‘‘hot center’’ 16 having a coordination structure close to that of Yb in the stable fcc phase, which initiates a self-sustaining pro- cess of avalanche crystallization. Pure amorphous films of Bi and Be are superconductors even during condensation, once they reach a certain thick- ϳ ness ( 5 nm). Heating the samples from 4.2 K to Tc in our FIG. 3. Reduced electrical resistance of amorphous vanadium films of thick- technique is done in approximately half an hour. By this time ness 12 nm ͑curve 1͒ and 20 nm ͑curve 2͒ as functions of the time of a hold a substantial part of the relaxation process will already have ϭ ͑ ͒ Ϫ Ϫ at Ti 4.2 K a andaplotofln͓(R0 Req)/(R Req)͔ versus t for these same films ͑b͒. occurred. For this reason, in studying relaxation processes it is important to take into account the relation between the time of the relaxation process and the time necessary for ϭ4.2 K, i.e., at a temperature even lower than the average measurement of the relaxing property. temperature of the film during its condensation. As is known However, in our experiments it is methodologically from studies of the structural relaxation of amorphous metal- simple to heat the films from 4.2 K to liquid hydrogen tem- lic alloys, at each temperature of an isothermal hold an in- perature (Ϸ20.4 K) over a time of the order of 1–2 min. vestigated property changes rapidly at first and then progres- This makes it possible to study at Tϭ20.4 K the relaxational sively slows in time, although in many cases there is no change of the resistance of amorphous films obtained at tem- discernable tendency for the property to stabilize at some peratures close to that of liquid helium. Of course, this is 3 Ͼ final value. This type of relaxation ordinarily takes place for possible only for thin films for which Tcr 20.4 K. Figure 4a annealing below the glass temperature Tg . A glass relaxes in shows the relaxational decrease of the resistance of an amor- the direction of the greatest equilibrium but, as a rule, does phous Bi film Ϸ8 nm thick, obtained by condensation of a not reach it. liquid-helium cooled substrate and rapidly heated to liquid Amorphous films of monoatomic metals behave in an hydrogen temperature. The time dependence of the resistance Ͻ ϭ analogous way during an isothermal hold at Ti Tcr , but in of this film during an isothermal hold at Ti 20.4 K is de- their case equilibrium is often attained. As an example, Fig. scribed satisfactorily by Eq. ͑1͒ with nϭ1/2 ͑see Fig. 4b͒.In 3a shows the relative change of the resistance of amorphous this case, as we see in Fig. 4a, ␦Ϸ0.74%. We note that the vanadium films at a temperature of 4.2 K as a function of the relaxational change of the resistance of a Bi film of approxi- hold time after condensation was stopped. mately the same thickness, condensed at a substrate tempera- As we see from Fig. 3a, the resistance R of the V films ture of Ϸ20.4 K, is only 0.08% during a 150-min hold at this decreases rapidly during the first 15 min of the hold, after same temperature. In the latter case, apparently, the relax- which R reaches a certain final value ͑at which ␦Ϸ0.4%͒. ation was practically completed during condensation. In many cases the relaxational change in the properties The same situation is observed for gallium. The resis- of amorphous solids is described satisfactorily by the tance of an amorphous Ga film Ϸ10 nm thick, condensed on 2 ϭ expression a liquid-hydrogen cooled substrate and held at Ti 20.4 K t n for 25 min after the evaporator was turned off, decreased by Ϫ ϭ͑ Ϫ ͒ ͩ Ϫ ͪ ͑ ͒ ͑ ͒ R Req R0 Req exp ␶ , 1 only 0.03%. During a further hold up to 50 min the resis- tance of this film remained unchanged. where R is the instantaneous value of the property ͑in our For amorphous manganese films an isothermal hold at ͒ ϭ case the resistance at time t, R0 and Req are the initial (t Ti 4.2 K for several hours did not lead to a noticeable ϭ0) and equilibrium (␶→ϱ) values of R;2 ␶ is the charac- change in resistance. However, after heating of the film to 838 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Kuz’menko and B. G. Lazarev

R/R0(T) in the case of a continuous heating of this film. The path of curve 2 shows that the relaxational decrease of the resistance of an amorphous Mn film during a 22-day hold at room temperature is equivalent to a continuous heating of that film at a rate of Ϸ3 K/min up to TϷ350 K. In the case of Mn the change of R with time during the isothermal hold cannot be described using a single characteristic relaxation time ␶, i.e., by a relation of the form ͑1͒. Relaxation phenomena are sometimes manifested most interestingly in the superconducting properties. It has been noted7,8,11 that in certain cases amorphous Bi and Ga films have a finite resistance immediately after condensation is stopped, and it decreases smoothly with hold time at 4.2 K until a transition occurs to the superconducting state. When the experimental conditions are changed so as to reduce to a minimum the introduction of impurities into the condensed layer of metal, this effect is eliminated,11 i.e., the supercon- ductivity arises right during the process of condensation. We studied the behavior of the resistance of amorphous Bi films containing up to Ϸ50 at.% hydrogen, during and after their condensation. Without exception, in all experi- ments the superconductivity of the films arose during con- densation, when the thickness of the film reached 5–6 nm.3͒ Apparently the aforementioned anomalous behavior of the resistance of amorphous Bi films7,8,11 during a hold at 4.2 K was due to some other impurities. In amorphous beryllium films stabilized by a hydrogen impurity we routinely observed a phenomenon similar to that described in Ref. 11. Pure beryllium films, upon reaching a FIG. 4. Relaxational decrease of the resistance of an amorphous Bi film ϳ ϭ ͑ ͒ ͓ Ϫ Ϫ ͔ 1/2 thickness of 5 nm, become superconducting during during a hold at Ti 20.4 K a ;aplotofln(R0 Req)/(R Req) versus t 16 for this same film ͑b͒. condensation. At a smaller thickness or if the film has been heated above Tc , the superconductivity sets in immediately after the evaporator is shut off ͑curve 1 in Fig. 6, where the room temperature, which is a process accompanied by an arrow indicates the time at which the condensation was irreversible decrease in resistance ͑see Figs. 2 and 5͒,an stopped͒. Ϸ isothermal hold at Ti 295 K leads to a further drop in resis- The parts of the curves to the left of the arrow show the tance. Figure 5 shows portions of the heating curves for two change in resistance of the films during condensation before identical amorphous Mn films. Curve 1 corresponds to a con- the evaporator was shut off. Curves 2 and 3 in Fig. 6 ͑to the tinuous heating at an average rate of Ϸ3 K/min, and curve 2 right of the arrow͒ describe the behavior of Be–H films after corresponds to heating with an isothermal hold at room tem- the evaporator was shut off. It is seen that the superconduc- perature for 22 days. tivity of the Be–H films at 4.2 K arises during the first The interpolation of curve 2 between points A and C minute or so of the hold at this temperature. After the film is ͑ ͒ Ϸ shows the presumed by analogy with curve 1 path of heated to T 13 K over a time of about half an hour, their Tc was already about 10 K with a transition width of Ϸ0.3 K. Ͻ Ͻ Thus in the time interval 1 min t 30 min the value of Tc of the Be–H films increases from below 4.2 to Ϸ10 K. Here the normal resistance of these films decreases by approximately 2% in comparison with the resistance at the time the conden- sation was stopped. Thus the relatively small relaxational decrease of the resistance of Be–H films is accompanied by a significant change in the superconducting transition tem- perature. It should be noted that curves 2 and 3 in Fig. 6 describe the behavior of two Be–H films of the same thick- ness and containing approximately the same amount of hy- drogen. One can discern a tendency toward increasing delay of the onset of superconductivity in the as-condensed films of Be–H with increasing hydrogen concentration. Unfortu- FIG. 5. Temperature dependence of the resistance during the heating of nately, we cannot determine the hydrogen concentration in amorphous Mn films condensed on a liquid-helium cooled substrate: 1—continuous heating at a rate of Ϸ3 K/min; 2—heating with an isother- these films to sufficient accuracy, and anyway it decreases ϭ mal hold at Ti 295 K for 22 days. rapidly as the samples are heated to room temperature. Very Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Kuz’menko and B. G. Lazarev 839

becomes higher ͑but long-range three-dimensional order does not arise͒. The distances over which the atoms move during relaxation is around 0.1 of the average interatomic distance, but the positions of almost all the atoms change.1 For example, the relaxational change of the resistance in an amorphous Mg70Zn30 alloy is accompanied by a displace- ment of the atoms by 0.01–0.03 nm.21 The most complete electron diffraction studies ͑using the method of constructing the radial distribution functions of the atomic density͒ of the structure of low-temperature con- densates of Bi, Ga, Yb, and Be have been done by Komnik and co-workers.9,10 In particular, it was shown that irrevers- ible changes in the electrical properties of amorphous Bi films of small thickness (Ͻ5 nm) with increasing tempera- ture ͑analogous to those seen in Fig. 1b͒ are due mainly to changes in the coordination structure. Structural transforma- tions in amorphous Bi films of thickness 40–50 nm have also Ϸ been observed to occur during a hold at Ti 4.2 K after con- Ͻ densation was stopped. On heating to Th Tcr the coordina- tion structure of the Bi films of all the thicknesses studied became approximately the same ͑‘‘typical’’͒. As the thick- ness of the amorphous ytterbium films was increased ͑from 2 to 20 nm͒ a decrease in the coordination number from 9 to 7 was observed. For Yb films of small thickness the coordina- tion number decreased from 9 to 7 as the temperature was raised.10 Qualitative results of electron diffraction studies of low- temperature condensates of Be have shown that with increas- ing temperature the changes of the radial distribution func- tion manifest the same tendencies as in the case of ytterbium and bismuth.10 Thus we can assume that the relaxational changes of the ϭ FIG. 6. Change in voltage with time at Ti 4.2 K for amorphous Be–H films of thickness Ϸ27 nm, after condensation was stopped ͑measurement resistance and Hall effect described above and shown in current Ϸ30 ␮A͒. Figs. 1–6 are due to the structural changes observed in films of this sort in Refs. 9 and 10. Ultimately these structural changes amount to a displacement of the metal atoms into ͑ ͒ rough estimates according to formula 16 from Ref. 19 give positions with lower free energy. It might seem that these values of the initial hydrogen concentration of 1–3 at.% in displacements are negligibly small, since even the crystalli- the Be–H films. zation of monoatomic amorphous metals is realized via a The superconductivity temperature of the amorphous Bi, diffusionless mechanism involving single ‘‘hops’’ of atoms Ͼ Ga, and Be–H films established after heating to Th Tc no across the crystallization front, and the length of such ͑ longer varies on further heating to very high but less than ‘‘hops’’ is a small fraction of the interatomic distance.16 ͒ Tcr temperatures, in contrast to the behavior of certain amor- It should be emphasized that, in contrast to the decrease phous metallic alloys, where, as a result of relaxation pro- in resistance as a result of relaxation of amorphous Bi films, Ͻ cesses on heating to Th Tcr , a lowering of Tc by 0.15–0.35 4,5,20 the process of crystallization of these films is accompanied K has been observed. We observed approximately the by a sharp increase in resistance ͑see Fig. 1a͒. In the case of ͑ ͒ same decrease in Tc by 0.15–0.20 K for Be–H films amorphous Mn films the relaxation process on heating is ͑ ͒ 50–70 nm thick heated to the temperature at which crys- characterized by an increase of the Hall constant at T Ϸ tallization begins ( 66– 68 K), when, according to our esti- ϭ4.2 K ͑Fig. 2b͒, whereas the crystallization of these films is mates, 1.5–2.5% of the sample had already formed the hcp ͑ ͒ accompanied by a decrease of RH at this same temperature phase of Be. The cause of the decrease in Tc in that case is to such a degree that it changes sign.18 Furthermore, unlike the proximity effect. the relaxation process ͑see Figs. 3 and 4͒ the initial period of isothermal crystallization of amorphous metals is character- 4. DISCUSSION OF THE RESULTS ized by acceleration of the change in resistance with hold Let us start by noting some of the most general trends time.16 These facts are evidence that the relaxation processes discerned in the study of the irreversible structural relaxation in the cases discussed above are in fact occurring in the of amorphous metallic alloys.1–3 Diffraction studies show amorphous phase of the metals and are not due to an initial that upon structural relaxation the height of the interference stage of crystallization. peaks increases and the dips become more individual, i.e., The main relaxation process of monoatomic amorphous the degree of regularity of the short-range structural order metals ͑V, Yb, Bi͒ is a simple exponential process with a 840 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Kuz’menko and B. G. Lazarev single relaxation time, since it is satisfactorily described by ϭ Ϫ ЈϷ Ϫ ͑ ͒ v f ¯v v0 ¯v 1.17v0 , 3 Eq. ͑1͒. Although the time is raised to a power different from Ј unity in that equation, its differentiation with respect to time where ¯v, v0 , and v0 are, respectively, the average volumes gives a linear differential equation of the first degree:2 per atom in the amorphous film, in the ideal close-packed crystal, and in the close-packed disordered structure ͑Bernal glass24͒, which models an amorphous solid with internal d Ϫ ͒ϭ Ϫ ␶Ϫn nϪ1͒ Ϫ ͒ ͑ ͒ equilibrium. The regions of free volume in amorphous alloys ͑R Req ͑ n t ͑R Req . 2 dt have been detected experimentally.25 Positron annihilation experiments have shown that the amorphous structure con- ␶ ͑ The characteristic relaxation time is determined by the tains vacancylike defects with a size of (0.2– 0.7)d0 d0 is Ϫ Ϫ ͒ slope of the linear plot of ln͓(R0 Req)/(R Req)͔ versus t or the size of an atom of the metal , the concentration of which t1/2 shown in Figs. 3b and 4b and has a value, according to reaches 3%.26 These defects also contribute to the free vol- these plots, of Ϸ5 min for the vanadium films and Ϸ33 min ume of amorphous films. The atomic rearrangement in an for the bismuth. The physical meaning of the relaxation time amorphous film is accompanied by a redistribution and es- is not always clear; there is apparently no simple relation cape of the free volume. Proof of this is given by the increase between ␶ and the characteristics of the material.2 It can be in the density of the amorphous alloys on annealing, which is assumed1 that ␶ is the time required for the atoms of the due to the escape of the free volume, the time dependence metal to form a new configuration when they have occupied being of a logarithmic character.27 positions close to equilibrium. It appears that the change of the topological structure of As can be seen from Table I and Figs. 3–5, the relax- amorphous films of Bi, Ga, Yb, and Be9,10 and the accompa- ational changes of the resistance of the investigated amor- nying change of the electronic properties, which is described phous films during an isothermal hold at a temperature Ti in the present paper, are also directly due to the redistribution and during continuous slow heating to a higher temperature of the free volume both during a hold of the films at a fixed Ͻ Ͻ Ͻ Th (Ti Th Tcr) are quantities of the same order. Appar- temperature and during their heating to Th Tcr . This view ently, during the relaxation of amorphous metals a certain is supported, in particular, by experiments on the structural equivalence of parameters ͑temperature and time͒ is mani- relaxation of ytterbium films deposited at a nonzero angle ␣ fested which is characteristic of temperature–time between the direction of the molecule beam and the normal superposition.2 This circumstance is in agreement with the to the substrate. It would seem that the relaxational change in aforementioned linearity of the main relaxation process in the resistance of such films should be larger than in the case the films under study. of films deposited normally, since the latter have a higher In amorphous manganese films there is apparently a co- density.28,29 In other words, the films deposited at an angle valent component of the bonding between atoms in addition have a larger free volume. Indeed, we found that a hold at ϭ to the metallic. That would account for the very high values Ti 4.2 K of amorphous Yb films condensed at an angle ␳Ϸ Ϯ ϫ Ϫ6 ⍀ ␣Ϸ ␦ of the resistivity ( (420 40) 10 •cm) and crystal- 50° leads to an increase of the parameter by an order of Ϸ 18 ͑ lization temperature on heating (Tcr 380 K). As we have magnitude in comparison with films deposited normally un- said, the relaxational decrease of the resistance in amorphous der otherwise similar conditions͒. Mn is not described by a single relaxation time. It seems that It is known that the insertion of hydrogen atoms into the in this case for describing the stabilization of the amorphous lattices of metals usually causes an increase in volume and a state it is necessary to take into account a spectrum of pos- distortion of the lattice. Nevertheless, the presence of up to sible relaxation times.2 The presence of such a spectrum can 50 at.% hydrogen in amorphous Yb and Bi films14,30 has reflect the objective participation of various molecular pro- little effect on their electronic characteristics, stability, and cesses. To explain these processes it will be necessary, in kinetic parameters of crystallization. Apparently, because of particular, to perfect techniques that will permit a rapid ap- the large size of the Yb and Bi atoms the micropores in their proach to the temperature of the isothermal hold in order to amorphous structures are larger in size than the hydrogen obtain more-precise data on the time dependence of the atoms, and the presence of the latter does not have much change in resistance at different temperatures. effect on the parameters of the short-range order. It can be The state of atomic configurations in as-condensed assumed that the low-temperature mobility of hydrogen at- amorphous films of pure metals is apparently close to the oms in the amorphous phases of Yb and Bi are high, possibly labile state predicted by Palatnikov and co-workers.22 There- because of the appearance of quantum effects ͑subbarrier fore the amorphous films of the metals studied ͑except for tunneling͒. For this reason the relaxation processes in amor- Mn͒, which were condensed at TϾ4.2 K, relax noticeably phous Bi–H and Yb–H films occur just as fast as in the pure even at a lower temperature, transforming into a metastable films. amorphous state with a lower free energy. The mechanism of In beryllium, whose atomic radius is 38–41% smaller local rearrangement of the atomic configurations by means than the atomic radii of Bi and Yb, hydrogen stabilizes the of structural relaxation are not yet sufficiently clear. The amorphous phase substantially.31 Nevertheless, even in most widely used description of the atomic transport and amorphous Be–H films the diffusive mobility of the hydro- structural relaxation is based on the free-volume theory.23 It gen atoms is evidently quite high. The activationless re- appears that in our experiments an important condition facili- grouping of hydrogen atoms in the as-condensed films of tating the atomic rearrangement is the presence in the amor- Be–H is, we believe, responsible for the delay in the onset of phous films of a ‘‘frozen’’ free volume v f , which is defined superconductivity in them, since in pure Be films such a by Turnbull and Cohen23 as delay is not observed ͑see curve 1 in Fig. 6͒. Finally, the Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Kuz’menko and B. G. Lazarev 841 behavior of curves 2 and 3 in Fig. 6 is not the result of a ͑e.g., Mn͒, for which a description of the stabilization re- simple cooling of the film after condensation is stopped. A quires taking into account a spectrum of possible relaxation study of the thermal regime during avalanche crystallization times. of amorphous Yb films32 has shown that under identical ex- 5. It appears that the process of structural relaxation of perimental conditions the cooling of the avalanche crystalli- monoatomic amorphous metals is due to a redistribution of zation front from a temperature Ϸ150 K16,17 to 5.6 K occurs the free volume in the samples. in Ϸ10 ms, i.e., the rate of cooling of the superheated film 6. In amorphous metal–hydrogen films the redistribution under these conditions reaches ϳ15000 K/s. of the free volume is apparently combined with an activa- Since amorphous beryllium is a superconductor with tionless regrouping ͑subbarrier tunneling͒ of hydrogen at- weak coupling, one can use the BCS formula for Tc : oms. 7. Comparatively small relaxational changes in the resis- 1 Ϸ ⌰ ͩ Ϫ ͪ ͑ ͒ tance in amorphous Be–H films ͑at T ϭ4.2 K͒ are accompa- Tc 1.14 D exp ͑ ͒ , 4 i N EF V* nied by a substantial change in the superconducting proper- ⌰ where D is the Debye temperature, N(EF) is the density of ties, a fact which we attribute to enhancement of the states of the electrons at the Fermi level, and V* describes electron–phonon interaction in the process of structural re- the interaction between two electrons through the exchange laxation of these films. of phonons with allowance for the screened Coulomb inter- action. * As we have said, the relaxation of the as-condensed E-mail: [email protected] †Deceased. Be–H films, leading to an increase of T from 4.2 to Ϸ10 K, 1͒ р 16 c For amorphous V films the value of Tc 3K. is accompanied by a decrease of the electrical resistance ͑in 2͒If the equilibrium value of the resistance was not reached during the iso- Ϫ1 the normal state͒ by Ϸ2%. The accompanying change in the thermal hold, then Req was determined by extrapolating the R versus t Debye temperature is hardly noticeable. For example, the plot to t→ϱ. 3͒ It is known that the value of Tc for amorphous metallic films decreases complete structural relaxation of some amorphous alloys 13,16 ͑ 5 ⌰ strongly with decreasing thickness below 10 nm. e.g., Zr54Cu46) leads to a decrease of D by only 2–3%. Elementary estimates using the formulas of free-electron theory show that the observed Ϸ2% decrease in the resis- tance of amorphous Be–H films corresponds to an increase 1 K. Suzuki, H. Fujimori, and K. Hasimoto, Amorphous Metals ͓Russian in the density of states by Ϸ1%. ͑Here we have made the translation͔, Metallurgiya, Moscow ͑1987͒. 2 assumption that the mean free path of the electrons for these D. R. Ullman and R. W. Hopper, in Metallic Glasses J. J. Gilman and H. J. Leamy ͑Eds.͒, ASM International, Metals Park, Ohio ͑1979͒; Metal- films, which is approximately equal to the interatomic lurgiya, Moscow ͑1984͒. 16 distance, does not change in the course of the relaxation.͒ 3 A. L. Greer, J. Non-Cryst. Solids 61–62,787͑1984͒. ͑ ͒ 4 According to formula 4 , a change of the value of Tc Z. Altounian, Tu Guo-hua, and J. O. Strom-Olsen, Solid State Commun. 40,221͑1981͒. from 4.2 to 10 K corresponds to an increase in N(EF)V* by 5 Ϸ P. Garoche, Y. Calvayrac, W. Cheng, and J. J. Veyssic, J. Phys. F.: Metal. 18.5%. Apparently, the sharp increase in Tc as a result of Phys. 12, 2783 ͑1982͒. relaxation of the as-condensed Be–H films is due mainly to 6 J. Hillairet, E. Balanzat, N.-E. Derradji, and A. Chamberod, J. Non-Cryst. growth of the electron–phonon interaction parameter V*. Solids 61–62, 781 ͑1984͒. 7 The same conclusion was reached by the authors of Ref. 4, B. G. Lazarev, E. E. Semenenko, and A. I. Sudovtsov, Fiz. Met. Metall- oved. 28, 954 ͑1969͒. who showed that the relaxational change in Tc on heating of 8 V. V. Postnikov, Author’s Abstract of Candidate’s Dissertation ͓in Rus- amorphous Cu–Zr films is due specifically to a change of the sian͔, Voronezh ͑1971͒. parameter V* and not of N(E ), since, according to their 9 Yu. F. Komnik, B. I. Belevtsev, and L. A. Yatsuk, Zh. E´ ksp. Teor. Fiz. 63, F ͑ ͓͒ ͑ ͔͒ magnetic susceptibility measurements, the change in N(E ) 2226 1972 Sov. Phys. JETP 36,1177 1973 . F 10 Yu. F. Komnik and L. A. Yatsuk, in Whisker Crystals and Thin Films ͓in is small. Unfortunately, we cannot propose a more rigorous Russian͔, Part 2, Voronezh ͑1975͒,p.85. 11 explanation for the observed changes of Tc with time in E. E. Semenenko, B. G. Lazarev, and V. I. Tutov, Abstracts of the 21st All-Union Conference on Low Temperature Physics ͓in Russian͔, Part 1, amorphous Be–H films, since an estimate of Tc requires de- Kharkov ͑1980͒, p. 104. tailed knowledge of the electronic and phonon spectra, and 12 33 V. M. Kuz’menko, B. G. Lazarev, V. I. Mel’nikov, and A. I. Sudovtsov, hydrogen is known to have a strong influence on both. Ukr. Fiz. Zh. ͑Russ. Ed.͒ 21, 883 ͑1976͒. 13 Yu. F. Komnik, Physics of Metallic Films ͓in Russian͔, Atomizdat, Mos- cow ͑1979͒. 14 V. M. Kuz’menko and A. N. Vladychkin, Fiz. Tverd. Tela ͑St. Petersburg͒ CONCLUSIONS 41,177͑1999͓͒Phys. Solid State 41,155͑1999͔͒. 15 J. C. Suits, Phys. Rev. 131, 588 ͑1963͒. 1. In amorphous films of monoatomic metals there is 16 V. M. Kuz’menko, Doctoral Dissertation ͓in Russian͔, B. Verkin Institute appreciable relaxation of the electrical resistance even at a for Low Temperature Physics and Engineering, National Academy of Sci- temperature of 4.2 K, due to a structural rearrangement. ences of Ukraine, Kharkov ͑1992͒. 17 ͑ ͒ 2. The structural relaxation of amorphous metals is not V. A. ShklovskiŽ and V. M. Kuz’menko, Usp. Fiz. Nauk 157,311 1989 . 18 V. M. Kuz’menko and V. I. Mel’nikov, Fiz. Met. Metalloved. 50,984 an incipient crystallization process but is a process of stabi- ͑1980͒. lization of the amorphous structure. 19 H. L. Caswell, in Physics of Thin Films, edited by G. Hass, Vol. 1 ͓Aca- 3. This relaxation, as a rule, is a simple exponential pro- demic Press, New York ͑1963͒; Mir, Moscow ͑1967͔͒. 20 cess with a single characteristic relaxation time. P. Esquinazi, M. E. De la Cru´z, and F. De la Cru´z, Physica B & C 108, 1215 ͑1981͒. 4. An apparent exception is the case of amorphous met- 21 T. Mizoguchi, H. Narumi, N. Akutsu, N. Watanable, N. Shiotati, and als with a substantial covalent contribution to the bonding M. Ito, J. Non-Cryst. Solids 61–62, 285 ͑1984͒. 842 Low Temp. Phys. 27 (99–10), September–October 2001 V. M. Kuz’menko and B. G. Lazarev

22 28 ͑ ͒ L. S. Palatnik, Yu. A. BykovskiŽ, P. A. Panchekha, A. G. Dudoladov, V. I. L. S. Palatnik and A. I. Fedorenko, Fiz. Tverd. Tela Leningrad 7,3163 Verchenko, and S. V. Marun’ko, Dokl. Akad. Nauk SSSR 254, 632 ͑1980͒ ͑1965͓͒Sov. Phys. Solid State 7, 2561 ͑1965͔͒. ͓Sov. Phys. Dokl. 25,770͑1980͔͒. 29 N. G. Nakhodkin and A. I. Shaldervan, Thin Solid Films 10,109͑1972͒. 23 D. Turnbull and M. H. Cohen, J. Chem. Phys. 52, 3038 ͑1970͒. 30 B. I. Belevtsev, Yu. F. Komnik, V. I. Odnokozov, and A. V. Fomin, Fiz. 24 J. D. Bernal, Proc. R. Soc. London, Ser. A 280, 299 ͑1964͒. Nizk. Temp. 8, 549 ͑1982͓͒Sov. J. Low Temp. Phys. 8,272͑1982͔͒. 25 T. Egami, in Glassy Metals,H.J.Gu¨ntherodt and H. Beck ͑Eds.͒ 31 V. M. Kuz’menko, V. I. Mel’nikov, T. P. Chernyaeva, and V. V. Bryk, Fiz. ͓Springer-Verlag, Berlin ͑1981͒; Mir, Moscow ͑1983͔͒,p.45. Met. Metalloved. 8,41͑1990͒. 26 Yu. V. Barmin and I. V. Zolotukhin, in Physics and Physical Chemistry of 32 V. M. Kuz’menko and V. I. Mel’nikov, Fiz. Nizk. Temp. 14,972͑1988͒ Amorphous (Glassy) Metallic Materials ͓in Russian͔, Nauka, Moscow ͓Sov. J. Low Temp. Phys. 14,533͑1988͔͒. ͑1987͒,p.80. 33 ͓ ͔ R. A. AndrievskiŽ, Materials Science of Hydrides in Russian , Metal- 27 ͑ ͒ T. N. MikhaŽlova, V. V. Molokanov, and G. A. Kuvshinov, in Physical lurgiya, Moscow 1986 . Chemistry of Amorphous (Glassy) Metallic Materials ͓in Russian͔, Nauka, Moscow ͑1987͒,p.127. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Isotope effects and the manifestation of 2D phase transitions in the kinetics of low- temperature „down to 5 K… hydrogen adsorption * ˘ V. D. Osovski , Yu. G. Ptushinski , V. G. Sukretnyi, and B. A. Chuıkov Institute of Physics of the National Academy of Sciences of Ukraine, pr. Nauki 46, 03650 Kiev, Ukraine ͑Submitted April 2, 2001͒ Fiz. Nizk. Temp. 27, 1138–1147 ͑September–October 2001͒ The adsorption–desorption characteristics of hydrogen isotopes on the ͑110͒ surfaces of W and ϳ Mo are investigated at a temperature Ts 5 K. It is found that the characteristics for ͑ ͒ ϳ hydrogen and deuterium are very different. Isothermal desorption from W 110 at Ts 5Kis observed in the case of H2, while that of D2 is practically absent. The desorption of H2 is suppressed as the flux of molecules to the surface increases; this, in the opinion of the authors, is a consequence of phase transition from a 2D gas to a 2D condensate. The initial sticking ͑ ͒ probability of H2 on W 110 depends strongly on the flux, while that of D2 is nearly independent of flux; in the H2 case this is explained by the formation of nuclei of the 2D condensed phase during the lifetime of the molecule in the precursor state. In the case of H2 adsorption on Mo͑110͒ a sharp and deep minimum is observed in the dependence of the sticking probability on the coverage, which is absent in the case of D2. These isotope effects are explained by the deeper position of the levels of the zero-point vibrations of the heavier molecule ͓ ͓ ͔ D2.©2001 American Institute of Physics. DOI: DOI: 10.1063/1.1414574

1. INTRODUCTION In view of what we have said, we undertook an experi- mental investigation of the low-temperature adsorption of the ͑ ͒ The study of the adsorption interaction of hydrogen with hydrogen isotopes H2 and D2 on the 110 surfaces of single the surface of metals is of interest from at least two stand- crystals of tungsten and molybdenum. Adsorption systems points. First, hydrogen is a participant in such important based on tungsten and molybdenum have become wide- catalytic reactions as the synthesis of ammonia and is also a spread in surface physics as model systems with a compara- prospective environmentally clean fuel. In the latter applica- tively easily obtained atomically clean surface. tion it is necessary to dissolve and store the hydrogen in a solid-state carrier. In both catalysis and dissolution the key 2. TECHNIQUES stage of the processes is the dissociative adsorption of hy- These studies were carried out in an ultrahigh-vacuum drogen. Second, in the case of hydrogen adsorption one can apparatus of the ‘‘black chamber’’ type, the general features expect significant isotope effects, especially at low tempera- of which are described in Refs. 6 and 7. The apparatus has tures. This is one of the reasons that have prompted us to do provisions for the formation of a molecular beam of the ef- an experimental study of the kinetics of H2 and D2 adsorp- fusion type, the straight-line registration of desorbed par- tion at a substrate temperature T ϳ5K. ϳ s ticles, and cooling of the sample to a temperature Ts 5K. The second reason was the desire to investigate the Here we give only a block diagram of the layout of the main weakly bound states of molecular adsorption of hydrogen in elements of the apparatus ͑Fig. 1͒. The vacuum conditions in connection with their possible participation in the mecha- the apparatus are extremely perfect: with the sample manipu- nism of dissociative adsorption in the capacity of precursor lator flooded with liquid helium and the hydrogen molecular states ͑prestates͒. The idea of the participation of prestates in beam open, the residual gas pressure is less than 10Ϫ11 Torr, the adsorption process has become widely disseminated. Two and the flux of molecules from the molecular beam source to types of prestates are distinguished: intrinsic and extrinsic. the sample is 4 orders of magnitude greater than the back- The first is realized above an unoccupied site and influences ground flux. ͑ ␪→ ͒ the initial sticking probability S0 at a coverage 0 , and Tungsten or molybdenum samples in the form of disks the second is above an occupied adsorption site with the 12 mm in diameter and 0.8 mm thick were welded on as the possibility of migration to an unoccupied site. The very small bottom of a tubular rocking manipulator. The samples were binding energy and, hence, the short lifetime of a molecule subjected to heat treatment in vacuum by the standard tech- on the surface at ordinary temperatures are characteristic nology for ensuring an atomically clean surface. The heating properties of prestates. Therefore, in order to stabilize and of the samples to temperatures TϽ1000 K was done by study the prestates it is necessary to cool the sample to means of the thermal radiation from an incandescent tung- liquid-helium temperature. Although the study of hydrogen sten coil, and to temperatures TϾ1000 K by electron bom- adsorption on metals at ordinary temperatures has been the bardment of the back side of the sample. The temperature subject of an enormous number of published works ͑see, e.g., was monitored by means of a tungsten–rhenium thermo- the reviews1–4͒, the number of experimental studies on low- couple. Three types of experiments were done: ϳ temperature adsorption of hydrogen is extremely limited. 1. On a substrate cooled to Ts 5 K a layer of hydrogen References to the papers we know of can be found in Ref. 5. ͑deuterium͒ was adsorbed until saturation, i.e., until a dy-

1063-777X/2001/27(9–10)/7/$20.00843 © 2001 American Institute of Physics 844 Low Temp. Phys. 27 (99–10), September–October 2001 Osovski et al.

FIG. 2. Isothermal desorption of H2 molecules: low flux during formation of the adlayer ͑1͒, high flux ͑2͒. FIG. 1. Diagram of the arrangement of the basic elements of the apparatus: 1—sample, 2—mass-spectrometric detector, 3—molecular beam source, 4—shield. 3. RESULTS AND DISCUSSION

3.1. Isothermal desorption of hydrogen from the W„110… surface and the influence of the molecular flux namic equilibrium was established between the incident and The isothermal desorption of hydrogen after the molecu- desorbing fluxes. During the formation of the adsorbed layer lar beam was shut off was investigated for several different ͑adlayer͒ the intensity of the beam of desorbing ͑and re- values of the flux to the surface during formation of the flected͒ molecules was continuously measured. These mea- adlayer.9 Figure 2 shows the time dependence of the rate of surements made it possible to determine the sticking prob- isothermal desorption of hydrogen molecules from adlayers ability S and the surface density n from the following formed at fluxes of ϳ1ϫ1013 and ϳ2ϫ1014 expressions: molecules/͑cm2s͒. Since the duration of the desorption pro- cess is considerably different for the different fluxes, in Fig. ͒ϭ Ϫ ͒ ͑ ͒ 2 the time is plotted on a logarithmic scale. It is seen that S͑t 1 I͑t /Im , 1 increasing the flux causes a sharp suppression of the isother-

t mal desorption. The number of molecules desorbed in the n͑t͒ϭF ͵ S͑t͒dt, ͑2͒ case of the high flux was smaller by a factor of 50 than for 0 the low flux. Let us begin with a discussion of the data for the case of where I(t) and Im are the ion currents of the detector at the a low flux during formation of the adlayer. The desorption time t and after the formation of a saturated adlayer, respec- process is conveniently divided into a rapid and a slow stage. tively, and F is the flux of molecules to the surface of the The rapid stage is completed in several seconds, and during ␪ ϭ sample. The coverage (t) n(t)/na , where na is the den- this stage less than 1% of the total number of molecules is Ϫ sity of surface atoms of the substrate (ϳ1.4ϫ1015 cm 2) for desorbed. We think that the rapid stage represents the desorp- the investigated faces of W and Mo͒. The intensity of the tion of a small number of molecules that had been in dy- molecular beam was calibrated by a method described in namic equilibrium when the flux was on, residing in a state Ref. 8. of the polylayer condensation type. The slow stage continues 2. After the formation of a saturated adlayer the molecu- for hundreds of seconds and, in our opinion, represents the lar beam was shut off and the isothermal desorption of mol- desorption from a state of physical adsorption ͑physisorp- ϳ ecules at Ts 5 K was observed. tion͒. 3. The saturated adlayer was desorbed as the temperature For analysis of the slow stage of isothermal desorption of the sample was gradually raised, and the thermodesorption we use the Polyani–Wigner equation in logarithmic form: spectrum, which characterizes the set of adsorption states, log͑Ϫdn/dt͒ϭlog ␯ ϩa log nϪ0.43E /kT , ͑3͒ was measured. d d s ␯ The isotope effects are most clearly manifested in ex- where n is the surface density of adsorbed molecules, d is periments of types 1 and 2, and the discussion below will be the frequency factor, a is the order of the desorption reaction, mainly devoted to a description of them. We will turn to the and Ed is the activation energy for desorption. As we see thermodesorption spectra in connection with the interpreta- from Eq. ͑3͒, the dependence of log(Ϫdn/dt)onlogn is lin- tion of the results. ear, and its slope can be used to determine the value of a. Low Temp. Phys. 27 (99–10), September–October 2001 Osovski et al. 845

FIG. 3. Dependence of the rate of isothermal desorption of H on the sur- FIG. 4. Thermodesorption spectrum in the case of a low flux of H2 mol- 2 ϳ face density. ecules. The rate of increase of the sample temperature was 0.3 K/s in the temperature interval 5–100 K, and ϳ7 K/s in the interval 200–700 K.

Figure 3 shows the experimentally measured dependence. It where ␶ ϭh/kT is the vibrational period of the molecules. can be approximated by two linear segments AB and CD. 0 s An estimate gives E Ϸ15 meV for the 6 K state and E For segment AB the value of a is close to unity, and for d d Ϸ25 meV for the 10 K state. On the basis of the results segment CD it is close to zero. These results indicate that the discussed above, we propose the hypothetical model of a isothermal desorption of physisorbed molecules occurs from hydrogen adlayer at T ϳ5 K which is illustrated in Fig. 5. two different states. We assume that segment AB represents s Let us now turn to a discussion of the data on the iso- desorption from a two-dimensional ͑2D͒ gas state, and seg- thermal desorption in the case of a high flux of hydrogen ment CD represents desorption with the participation of is- molecules during formation of the adlayer. If our assumption lands of a 2D condensed phase. Under conditions such that as to the influence of the islands of 2D condensate on the the 2D gas phase is exhausted and the rate of desorption isothermal desorption is correct, we would expect that with becomes very low, a substantial role is played by the mecha- increasing flux the probability of island formation would in- nism wherein the gas phase is replenished through 2D evapo- crease and, accordingly, so would the fraction of the surface ration of the islands of the 2D condensate. After a certain covered with the 2D condensed phase. It is known that the time this stabilizes the surface density in the 2D gas phase critical size of the nuclei of the condensed phase should de- ͑although the total number of adsorbed molecules is continu- crease with increasing supersaturation, i.e., with increasing ously decreasing͒, and the order of the desorption reaction flux of molecules. Then it becomes more probable for a gas– approaches zero. It is expected that the activation energy for condensate phase transition to occur during the time of the desorption from the 2D condensate into vacuum should be experiment. As we see in Fig. 2, increasing the flux did in- greater than from the 2D gas because of the attraction be- deed lead to strong suppression of the isothermal desorption. tween molecules in the condensate. We assume that the suppression of the isothermal desorption That the physisorbed H layer has a two-phase structure 2 at T ϳ5 K in the case of a high flux is due to the fact that a is also indicated by the thermodesorption spectrum measured s large fraction of the H molecules is adsorbed in the form of some time after the start of the slow stage of isothermal 2 a 2D condensed phase, whereas for the low flux only a small desorption ͑Fig. 4͒. Although the whole spectrum is shown, fraction of the adsorbed molecules is contained in the is- including the chemisorbed atomic phase ͑410 and 550 K͒, lands. The substantial increase in the 10 K peak in the ther- we are now interested only in the low-temperature part. It is reasonable to assume that the peak at 6 K corresponds to thermodesorption from the 2D gas state, and the peak at 10 K to that from the 2D condensate. We shall refer to the adsorption states by the temperature of the desorption peak. Assuming that at the temperature of the desorption peak, the lifetime ␶ of a molecule in the corresponding adsorption state is close to 1 s, we can estimate the activation energy for FIG. 5. Hypothetical model of an H2 adlayer at a sample temperature Ts ϳ desorption Ed from the Frenkel equation: 5K:1—chemisorbed monolayer of atoms; 2—physisorbed monolayer of molecules in the 2D gas and 2D condensate states; 3—molecules in a state ␶ϭ␶ ͒ ͑ ͒ 0 exp͑Ed /kTs , 4 of the polylayer condensation type. 846 Low Temp. Phys. 27 (99–10), September–October 2001 Osovski et al.

FIG. 6. Thermodesorption spectrum in the case of a high flux of H2 mol- ecules. The rate of increase of the temperature was ϳ1 K/s.

FIG. 8. Dependence of the sticking probability of H2 on the coverage for ϭ ϫ 14 ͑ ͒ ϫ 13 ͑ ͒ ϫ 12 ͑ ͒ ϫ 12 ͑ ͒ ͑ 2 ͒ modesorption spectrum in the case of a high flux ͑Fig. 6͒ F 2 10 1 ,4 10 2 ,9 10 3 , and 5 10 4 molecules/ cm s ; T is the gas temperature. argues in favor of the proposed interpretation. g If a change of state of the adsorbed molecules does not occur upon the change in flux, then the dynamic-equilibrium ␪ ͑ ͒ coverage s should follow the Langmuir isotherm. In our In Eq. 5 the pressure has been replaced by the flux. We see ␪ experiments we determined s from the number of isother- that at moderate values of the molecular flux the experimen- mally desorbed H2 molecules; the experimental results are tal data are satisfactorily described by the Langmuir iso- shown by the small squares in Fig. 7, and the continuous therm, but for FϾ1014 molecules/͑cm2s͒ a sharp deviation is curve is the Langmuir isotherm:10 observed, the result, we believe, of a phase transition from the 2D gas to the 2D condensate. ␪ ϭ ͓ ␯ ͑Ϫ ͒ϩ ͔ ͑ ͒ s F/ na d exp Ed /kTs F . 5 In experiments with deuterium, shutting off the molecu- lar beam after the formation of a saturated adlayer at Ts ϳ5 K led to a sharp drop in the ion current of the detector, i.e., the isothermal desorption of deuterium under these con- ditions is practically absent. The reason for this difference in behavior of the adsorbed D2 molecules from that of H2 mol- ecules may be found in the difference of the quantum prop- erties of these molecules. The level of the zero-point vibra- tions of the two-times heavier D2 molecule lies considerably deeper in the physisorption potential well than does that of the H2 molecule. Therefore, the activation energy for desorp- ϳ tion of the D2 molecule is greater, and at Ts 5 K no appre- ciable isothermal desorption of deuterium is observed.

3.2. Kinetics of hydrogen adsorption and the influence of the molecular flux In view of the results of the previous subsection, we expect, starting from the principle of detailed balance, that a change in the flux of molecules to the surface will also lead to a change in the adsorption kinetics. We measured the de- pendence of the sticking probability of the hydrogen mol- ecules on the W͑110͒ surface as a function of coverage, S(␪), for fluxes in the range 1012–1014 molecules/͑cm2s͒ ͑see Fig. 8͒.11 As we see from Fig. 8, increasing the flux FIG. 7. Langmuir isotherm; ᭿—experimental points. leads to significant growth of the initial sticking probability Low Temp. Phys. 27 (99–10), September–October 2001 Osovski et al. 847

͑ ␪→ ͒ ͑ ͒ ␪ S0 at 0 and of the maximum saturation coverage s , and also leads to the appearance and growth of a peak on the S(␪) curves. Let us discuss the possible mechanisms by which the ␪ molecular flux can influence S( ). An increase of S0 can arise if there is an interaction of molecules being adsorbed in an intrinsic prestate. Although we are talking about the initial sticking probability, which, strictly speaking, pertains to the clean surface, in the real experiment its measured value corresponds to some finite coverage of the surface. The time constant of our apparatus is 0.1 s, and during that time, at a flux of 5ϫ1012 molecules/͑cm2s͒, some 5ϫ1011 molecules/cm2 strike the surface, or, with a sticking probabil- ity ϳ0.1, a coating of 5ϫ1010 molecules/cm2 is formed. The average distance between molecules in this case is ϳ5 ϫ10Ϫ6 cm. In order to assess whether molecules found in the pr- estate, by moving along the surface, encounter and interact with each other, we can compare the distance between mol- ϳ ecules with the diffusion path length L at Ts 5 K. We have

not found any published data on the diffusion coefficient of FIG. 9. One-dimensional potential diagram: Ea and Ed are the activation ͑ ͒ energies of the transition to the states of chemisorption and desorption, H2 molecules along the W 110 surface, but in Ref. 12 there is information about the activation energy for the diffusion of respectively; Z p is the level of the zero-point vibrations. molecular hydrogen. Using those data, we can obtain a rough estimate of the diffusion coefficient from the expression a dynamic equilibrium density is established, with equal ϭ Ϫ ͒ ͑ ͒ D D0 exp͑ Em /kTs , 6 rates of adsorption and desorption. As to the cause of the increase of the sticking probability with increasing coverage where E is the activation energy for surface diffusion, and m upon the formation of a molecular adsorption phase, we hold the pre-exponential factor D was assumed equal to b2kT /h 0 s to the view expressed in one of our previous papers.16 We ͑b is the lattice constant of the substrate͒.13 The estimate suppose that the main cause of the growth of S is an increase gave a value DϷ10Ϫ7 cm2/s and LϭͱDtϭ10Ϫ4 cm. Thus L in the efficiency of a loss of kinetic energy ͑accomodation͒ is much larger than the average distance between molecules by the incident molecules in collisions with weakly bound even at the minimal flux, and it is quite probable that the molecules adsorbed previously. molecules encounter each other and interact in a time of Figure 10 shows the S(␪) curves for the adsorption of 0.1 s. Let us now discuss what specific mechanisms could lead to an increase in the initial sticking probability. If the elasti- cally reflected molecules, are not taken into account, the sticking probability can be expressed as follows:14,15 ␯ E ϪE Ϫ1 ϭͫ ϩ d ͩ a d ͪͬ ͑ ͒ S0 1 ␯ exp , 7 a kTs ␯ ␯ where a and d are frequency factors, and Ea and Ed are the activation energies for the transition from an intrinsic pr- estate to a state of chemisorption and desorption, respec- tively. What we have said is illustrated in Fig. 9 by a one- dimensional potential diagram. It is seen from Eq. ͑7͒ that an increase in the activation energy for desorption leads to an increase of S0 , and if the interaction of molecules in the prestate increases Ed , then this should be accompanied by an increase in S0 . It seems to us that the results considered in Sec. 3.1 are evidence that the increase in the initial sticking probability of hydrogen molecules with increasing flux is indeed caused by an increase in Ed . We see that increasing the flux suppresses the isothermal desorption of the hydrogen molecules, i.e., increases Ed . As to the specific mechanism for the increase in Ed , we propose that it is due to the for- mation of nuclei of the 2D condensed phase. The dependence of the maximum coverage on the flux of molecules ͑Fig. 8͒ is typical for an adsorption phase that is FIG. 10. Dependence of the sticking probability of D2 on the coverage for not maintained steady at the given temperature but for which Fϭ2ϫ1014 ͑1͒,4ϫ1013 ͑2͒, and 8ϫ1012 ͑3͒ molecules/͑cm2s͒. 848 Low Temp. Phys. 27 (99–10), September–October 2001 Osovski et al. deuterium. We see that these curves are qualitatively differ- ent from those for hydrogen. The maximum coverage and the initial sticking probability respond weakly to an increase in ͑ the flux of molecules during formation of the adlayer S0 even decreases somewhat͒. The maximum on the S(␪) curves is observed even at the lowest value of the flux, while it is absent in the corresponding case for H2. We assume that the differences in the adsorption kinetics of H2 and D2 and the influence of the flux on the adsorption kinetics are also determined by the difference in the positions of the levels of zero-point vibrations in the physisorption potential well. Since the level of zero-point vibrations of D2 is deeper, the activation energy for the desorption of D2 is ϳ substantially higher. Therefore, at Ts 5 K, where the phys- isorbed H2 layer is unstable, the physisorbed layer of D2 is stable and responds weakly to a change in flux. The reason for the absence of growth of S0 upon an increase in the flux of D2 molecules, we believe, is also the deeper position of the level of zero-point vibrations. The process of nucleation of the 2D condensed phase, which we think is responsible for the growth of the initial sticking probability for H2, does not take place in the case of D2 because of the deeper potential relief and, accordingly, the lower mobility of the molecules. Support for this view can be found in Ref. 12, where the ͑ ͒ FIG. 11. Initial sticking probability of H2 and D2 on Mo 110 versus the gas activation energy for surface diffusion of D2 is estimated as temperature. 6.6 meV, while for H2 it is 2.7 meV.

3.3. Adsorption of H and D on the Mo 110 surface 2 2 „ … estate. We note that a coverage ␪ϭ0.5 ͑in molecules͒ corre- It is of interest to compare the adsorption characteristics sponds to a monolayer of chemisorbed atoms. Thus S de- for hydrogen isotopes on the W͑110͒ surface with those for creases to an extremely small value when the atomic phase Mo͑110͒, which harldy differs in its atomic structure. In the of adsorption is saturated. The initially unfilled chemical case of Mo͑110͒ we measured S(␪) at a single value of the bonds of the surface atoms of Mo are saturated at the end of flux of molecules onto the surface, but we varied the gas the first stage and can no longer effect the capture and sub- temperature Tg in the molecular beam source and thus varied the kinetic energy of the incident molecules. The influence of a change in Tg on the adsorption characteristics of hydrogen on W͑110͒ was investigated by us previously.17 Figure 11 shows the dependence of the initial sticking ͑ ͒ probability for H2 and D2 on Mo 110 as a function of Tg .In both cases a monotonic decrease of S0 with increasing Tg is observed. This behavior of S0 argues in favor of an adsorp- tion mechanism involving an intrinsic prestate and the ab- sence of an appreciable activation barrier for transition to a state of chemisorption. In this respect the Mo͑110͒ surface is different from W͑110͒, for which a two-channel mechanism of H2 adsorption is observed: adsorption via a prestate, and 17 activational adsorption. The sticking probability of D2 is considerably larger than that of H2, a circumstance which is indicative of a phonon mechanism for the loss of kinetic energy of the molecules incident on the surface. On the other hand, the larger value of S0 for deuterium may also be due to the deeper-lying level of zero-point vibrations and, accord- ͑ ͒ ingly, to a larger value of Ed in Eq. 7 . Figure 12 shows the S(␪) curves for the adsorption of ͑ ͒ ϳ H2 and D2 on the Mo 110 surface for Tg 100 K. In the case of H2 the curve is extremely complex: after a slight increase of S there is a steep decrease to a sharp minimum at ␪ ϳ0.5, and then it passes through a maximum and finally falls gradually to zero. The behavior of S in the interval 0Ͻ␪ Ͻ ͑ ͒ 0.5 is evidence for adsorption through an extrinsic pr- FIG. 12. Sticking probability of H2 and D2 on Mo 110 versus the coverage. Low Temp. Phys. 27 (99–10), September–October 2001 Osovski et al. 849

sequent dissociation of H2 molecules, and the mass differ- of the level of zero-point vibrations of the heavier D2 mol- ence of the complex Mo/H and the H2 molecule is too large ecules in the physisorption potential well. ϳ for efficient accomodation. 2. At Ts 5 K one observes in the case of hydrogen, but In the interval 0.5Ͻ␪Ͻ1 the value of S increases not for deuterium, an increase in the initial sticking probabil- ͑ ͒ sharply owing to the formation of a weakly bound molecular ity S0 and in the maximum coverage of the W 110 surface phase, which facilitates the loss of kinetic energy of the in- when the flux of molecules to the surface is increased. We cident molecules. Granted, one cannot rule out the influence attribute this increase of S0 to an increased probability of of a ‘‘softening’’ of the surface phonons, as was conjectured formation of nuclei of the 2D condensed phase during the in Refs. 18 and 19. However, the increase in S is too sharp lifetime of the H2 molecules in the intrinsic precursor state. for it to be due to a simple accumulation of physisorbed H2 The absence of such an effect in the case of D2 adsorption is molecules. It is probable that a change of state of the phys- explained by the deeper position of the level of zero-point isorbed molecules occurs. We assume that the formation of vibrations in the physisorption potential well. nuclei of a 2D condensed phase occurs, which, as we have 3. The curve of the sticking probability of the H2 mol- said, suppresses the thermodesorption, further increasing S. ecule as a function of coverage for the Mo͑110͒ surface has The cresting and subsequent decline of S in the second stage a very sharp and deep minimum near ␪ϭ0.5. This feature is ␪ is probably due to the limited number of adsorption sites in absent on the S( ) curve for the adsorption of D2. We at- the physisorbed layer. tribute this difference in the adsorption kinetics of H2 and D2 Fig. 12 also shows the S(␪) curve for the adsorption of to the difference in the mobilities of these molecules along ͑ ͒ D2. This curve is qualitatively different from that for H2 in the Mo 110 surface, which, in turn, is due to the deeper ␪ϭ that there is no minimum near 0.5. To explain this differ- position of the level of zero-point vibrations of the D2 mol- ence we again turn to a consideration of the quantum prop- ecule. The very high mobility of the H2 molecules makes it erties of the H2 and D2 molecules. Because of the deeper possible to complete the formation of the atomic phase of position of the level of zero-point vibrations of the D2 mol- adsorption before the formation of the molecular phase be- ecule, its mobility in the extrinsic prestate should be substan- gins; this makes the surface inert in this stage of the process. tially lower than for H2. This circumstance favors the forma- We do not rule out the possibility of a contribution from tion of a molecular physisorbed phase of D2 even before the quantum diffusion of the H2 molecules in creating this high formation of the atomic phase is completed. The presence of mobility. The deeper potential relief for the D2 molecule and a weakly bound molecular phase leads to an increase in the its larger mass apparently make for a low probability of tun- sticking probability, compensating its decrease due to the neling diffusion for deuterium. saturation of the atomic phase. Because of their high mobil- This study was supported by the Foundation for Basic ity the H2 molecules easily ‘‘skip’’ to unoccupied sites, and Research of Ukraine, project No. 2.4/104. until those are all filled, the formation of a molecular phase of adsorption does not occur. It seems to us that one should *E-mail: [email protected] not rule out the possibility that quantum diffusion is partly responsible for the high mobility of the H2 molecules on the ͑ ͒ Mo 110 surface. 1 K. Christman, Surf. Sci. Rep. 9,1͑1988͒. As we showed in the previous subsection, in the case of 2 K. D. Rendulic, Surf. Sci. 272,34͑1992͒. ͑ ͒ 3 A. Gross, Surf. Sci. Rep. 32, 291 ͑1998͒. H2 adsorption on W 110 there is no sharp feature on the 4 ␪ ␪ϭ G.-J. Kroes, Prog. Surf. Sci. 60,1͑1999͒. S( ) curve near 0.5 like that considered above for 5 B. A. Chuikov, V. D. Osovskii, Yu. G. Ptushinskii, and V. G. Sukretnyi, H2 /Mo(110). This indicates that the potential relief along Surf. Sci. 448, L201 ͑2000͒. 6 the surface is substantially deeper for H2 /W(110) than for B. A. Chuikov, V. D. Osovskii, Yu. G. Ptushinskii, and V. G. Sukretnyi, H /Mo(110). This is also confirmed by the data of Ref. 12 Surf. Sci. 213, 359 ͑1989͒. 2 7 Yu. G. Ptushinskii, B. A. Chuikov, V. D. Osovskii, and V. G. Sukretnyi, on the relative values of the activation energies for surface Fiz. Nizk. Temp. 19, 570 ͑1993͓͒Low Temp. Phys. 19,406͑1993͔͒. ͑ ͒ ͑ ͒ 8 diffusion of H2 on W 110 and Mo 110 . Yu. G. Ptushinskii, B. A. Chuikov, V. D. Osovskii, and V. G. Sukretnyi, Ukr. Fiz. Zh. ͑Russ. Ed.͒ 44, 165 ͑1999͒. 9 4. CONCLUSION V. D. Osovskii, Yu. G. Ptushinskii, V. G. Sukretnyi, and B. A. Chuikov, JETP Lett. 67,959͑1998͒. 10 We have observed the following isotope effects in the E. A. Melvin-Hughes, Physical Chemistry, Izd. Inostr. Lit., Moscow ͑1962͒, p. 833. kinetics of low-temperature adsorption and desorption of hy- 11 B. A. Chuikov, V. D. Osovskii, Yu. G. Ptushinskii, and V. G. Sukretnyi, drogen. Surf. Sci. 473, 143 ͑2001͒. 1. Isothermal desorption of H molecules from the 12 P. P. Lutsishin, O. A. Panchenko, and V. F. Shpagin, Surf. Sci. 278,218 2 ͑ ͒ W͑110͒ surface at T ϳ5 K is observed. Increasing the flux of 1992 . s 13 R. Gomer, Rep. Prog. Phys. 57, 917 ͑1990͒. molecules onto the surface during the formation of a satu- 14 C. T. Rettner, H. Stein, and E. K. Schweizer, J. Chem. Phys. 89, 3337 ϳ ͑ ͒ rated adsorbed layer at Ts 5 K sharply suppresses the sub- 1988 . 15 Yu. G. Ptushinskii and B. A. Chuikov, Poverkhnost’ 9,5͑1992͒. sequent isothermal desorption of H2. The suppression of the 16 desorption is explained by the occurrence of a phase transi- V. D. Osovskii, Yu. G. Ptushinskii, V. G. Sukretnyi, and B. A. Chuikov, Fiz. Nizk. Temp. 23, 779 ͑1997͓͒Low Temp. Phys. 23,587͑1997͔͒. tion from the 2D gas to the 2D condensate. 17 V. D. Osovskii, Yu. G. Ptushinskii, V. G. Sukretnyi, and B. A. Chuikov, ͑ ͒ Under the same conditions, isothermal desorption of D2 JETP Lett. 60,586 1994 . 18 molecules was not observed. The reason for this difference in E. Hulpke and J. Ludechke, Phys. Rev. Lett. 68, 2846 ͑1992͒. 19 M. Okada, A. P. Baddorf, and D. M. Zehner, Surf. Sci. 373,145͑1997͒. the behavior of the H2 and D2 molecules, we believe, is the difference in their quantum properties—the deeper position Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Õ A vitrifying structure transition in the Dy Mo„112… adsorption system A. Fedorus,* V. Koval,† and A. Naumovets

Institute of Physics, National Academy of Sciences of Ukraine, 46 Nauki Ave., Kiev, 03028, Ukraine H. Pfnu¨r**

Institut fu¨r Festko¨rperphysik, Universita¨t Hannover, Appelstrasse 2, D-30167 Hannover, Germany ͑Submitted April 25, 2001͒ Fiz. Nizk. Temp. 27, 1148–1152 ͑September–October 2001͒ Annealing-driven irreversible structural transitions are studied by low-energy electron diffraction in submonolayer Dy films adsorbed on the Mo͑112͒ surface. In a wide coverage range ␪Ͼ0.07, Dy overlayers deposited at low temperatures (TϷ100 K) are ordered and keep their structure upon annealing up to 350–600 K. Near ␪ϭ0.68, the overlayers are stable to high-temperature ϭ annealing (Ta 1000 K) as well, whereas the denser films are metastable and transform to more stable ordered structures. An unusual annealing effect is found for ␪Ͻ0.58: the initially ordered metastable phases are replaced by phases having no extended order, which are assumed to be two-dimensional glasses. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1414575͔

INTRODUCTION RESULTS AND DISCUSSION The glassy state of matter has been attracting a great and To carry out this study, we have used the ultrahigh- abiding interest from both the basic and practical standpoints vacuum system and the experimental setup described ͑see, e.g., Ref. 1͒. There are a lot of methods of producing earlier.5–7 Briefly, a SPECTALEED instrument equipped various kinds of glasses. Among them, special attention is with a charge-coupled device ͑CCD͒ camera was used to paid at present to the preparation and study of thin glassy and study the surface structure. The stored video records of the amorphous films, which combine a number of useful proper- low energy electron diffraction ͑LEED͒ patterns were uti- ties provided by the disordered structure and low dimension- lized to measure the peak intensities for the particular spots. ality. In this communication, we report on a new method The contact potential difference method was applied for allowing one to obtain ultrathin ͑submonolayer͒ films which, monitoring the work function changes. The same molybde- in contrast to the usual amorphous films, are stable against num specimen as in our previous work6,7 was used for the annealing and retain their disordered structure in a broad ͑ ͒ temperature range. substrate. Its surface consists of 112 terraces as broad as Our results have been obtained with Dy submonolayers 350 Åϫ550 Å along the ͓¯1¯11͔ and ͓11¯ 0͔ axes, respec- on the Mo͑112͒ surface. The adsorbed Dy atoms possess an tively. The atomic beam of Dy was obtained from a piece of appreciable positive electrical charge, and their interaction dysprosium metal spot-welded to a tantalum ribbon serving with one another is of complex nature, including both dipole- as a heater. Auger electron spectroscopy was used to check dipole and substrate-mediated interactions. The energy of the cleanliness of the substrate and the adsorbate. The cov- such interactions can reach 10Ϫ2Ϫ10Ϫ1 eV for nearest erage ␪ was determined from the deposition time and work neighbors, so the adsorbed films are usually ordered below function measurements, which were calibrated by means of Tϭ100– 1000 K.2,3 Owing to the considerable dipole mo- LEED in the coverage range ␪ϭ0.68– 0.75. This range is ment associated with the adsorption bond ͑above1Datlow specific for the existence of a single-phase incommensurate Dy coverage͒, the adsorbate strongly reduces the work func- structure undergoing uniaxial compression and therefore is tion of the surface. The work function and structure of not likely to contain other phases which might be overlooked Dy/Mo overlayers have been studied earlier,4 but the by LEED. ϭ annealing-driven structural transitions and work function At annealing temperatures below Ta 350– 600 K, a set changes were not revealed. Specifically, the c(2ϫ2) lattice of ordered overlayer structures is found with increasing cov- was reported to exist up to ␪ϭ0.5, which is followed by erage from ␪ϭ0.07, until the formation of a close-packed uniaxially compressed lattices at ␪Ͼ0.5 in the submonolayer physical monolayer at ␪ϭ0.77. These structures in their ma- coverage range. The coverage is defined as the ratio of ad- jority are different from those reported in Ref. 4. As sorbate to substrate surface atom concentrations. We have we intend to discuss in detail the phase transitions for the investigated the structure of Dy adlayers both as-deposited at Dy/Mo͑112͒ system in a separate paper, we will not present a low temperature (Tϭ100 K) and successively annealed in here the whole series of ordered film structures observed in Ͻ Ͻ the range 100 K Ta 1200 K, with recording of the result- this work but will focus on the annealing-driven irreversible ing structures at the low temperature. The rate of temperature order-disorder transition. decrease after each annealing step did not exceed Ϫ5 K/s. In In general, the periodicity of an ordered film structure such experiments, we have found a strong effect of annealing differs from that of a substrate. Therefore LEED usually re- on the character of the order. veals an order in the adsorbed overlayer via extra ͑fractional-

1063-777X/2001/27(9–10)/4/$20.00850 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 Fedorus et al. 851

FIG. 3. Annealing-driven changes in intensity of fractional-order LEED beams specific to the structures observed at different Dy coverages. The beam intensity after annealing is always measured at Tϭ100 K.

FIG. 1. Low-energy electron diffraction patterns from Dy/Mo͑112͒ overlay- coverage. As a measure of the degree of order we use the ers. The pattern from the clean substrate ͑a͒ is presented for a reference, peak intensity of LEED. Above ␪Ͼ0.6, the annealing results with the principal axes indicated. Coverage ͑monolayers͒:0͑a͒; 0.25 ͑b͒; 0.68 ͑c͒; 0.45 ͑d͒. Annealing temperature, K: 200 ͑b͒; 800 ͑c, d͒. Electron in a usual effect of improvement of the order. This is illus- energy, eV: 67 ͑a,b͒:55͑c,d͒. trated in Fig. 3 for ␪ϭ0.68: the fractional-order beam inten- sity gradually increases and then levels off. The relevant LEED pattern from the eventually annealed overlayer is order͒ diffraction beams which are absent in the cases of a shown in Fig. 1c. The steady growth of the order, starting clean substrate or a disordered adsorbed layer. For example, from the lowest annealing temperature, is indicative of con- Fig. 1b shows a LEED pattern from an ordered (6ϫ1) Dy siderable mobility of adatoms already at 100 K. Designating overlayer which is observed in the coverage range ␪ as Tmob a certain minimum temperature at which adatoms ϭ0.07– 0.3. The pattern differs from that of the clean sub- can move over neighboring adsorption sites during the time strate ͑Fig. 1a͒ by the presence of five extra beams between Ͻ of an experiment, we conclude that Tmob is 100 K. Accord- ͓¯¯ ͔ the integer-order beams along the 111 direction. Figure 2 ing to this designation, Tmob can be expressed in terms of the ϫ Ϸ shows a possible model of the (6 1) structure which is parameters of surface diffusion as Tmob Ed/30kB , where Ed 2,8 consistent with the LEED pattern and the coverage value. It is the activation energy and kB is Boltzmann’s constant. should be mentioned that the (6ϫ1) structure is one of those Besides promoting adatom mobility, an increase of tempera- transforming into a glassy phase in the progress of annealing. ture can also cause an order–disorder transition if the tem- Patterns with fractional-order LEED spots are indeed ob- perature reaches some critical value Tc specified by the in- Ϸ ͑ ͒ served in the whole submonolayer coverage range at the low teraction energy of adatoms Ei as Tc Ei /kB see Ref. 8 . annealing temperatures mentioned above. A full list of the Ordering is realized within the time of an experiment ordered structures found will be given in the diagram of an- (10–102 s) if nealed states presented below. T ϽT ,orE Ͻ30 E ; ͑1͒ An increase of the annealing temperature exerts a differ- mob c d i ent effect on the order in Dy adfilms depending on their otherwise, upon the cooling that follows each annealing step, a frozen disorder is observed instead of order. The overlayer under consideration (␪ϭ0.68) is apparently what results when inequality ͑1͒ is fulfilled. It should be noted that the lattice retains its original symmetry under annealing only in a narrow coverage range close to ␪ϭ0.68. The denser films ͑from ␪ϭ0.68 up to the close-packed monolayer at ␪ ϭ0.77͒ are metastable and undergo irreversible transitions to the more stable ordered structures. In this case, LEED records the replacing of one pattern of superstructure spots by another. For the coverages 0.07Ͻ␪Ͻ0.58 the annealing results in an unusual effect: first only a limited improvement of the order occurs, which is followed by decay of the previous order without transition to any other ordered phase. This pe- culiarity is illustrated qualitatively by an irreversible disap- FIG. 2. A hypothetical hard-sphere model for an ordered dysprosium over- pearance of the fractional-order LEED features in pattern ͑d͒ ϫ ␪ϭ layer (6 1) corresponding to 0.3. The unfilled circles show the topmost of Fig. 1 and quantitatively by intensity curves with local substrate layer, and the gray disks show the adsorbate layer, with heavier ͑ ͒ darkness corresponding to greater vertical depth. The principal axes are maxima followed by virtual fading of intensity see Fig. 3 . Ͻ ͑ ͒ indicated. At low temperatures (Ta 200– 400 K), inequality 1 seems 852 Low Temp. Phys. 27 (99–10), September–October 2001 Fedorus et al.

FIG. 4. Diagram of annealing-driven changes in structure of Dy overlayer on Mo͑112͒. DO: disordered phase; 2D glass: two-dimensional glass ͑ ͒ ϫ ϫ ͑ ͒ shaded area ; r(nvar 2) and s(nvar 1): rectangular diatomic and skew- angle ͑primitive͒ structures, respectively, having lattice spacing varying with ␪; 2nd L: second layer; multi-L: multilayer structure. The other notations of FIG. 5. Coverage dependence of the work function changes obtained after Ͻ ordered structures are conventional. The arrows show the coverages at low- (Ta 350– 600 K) and higher-temperature annealing. which the specified structures form.

annealing-driven change of adatom mobility, which becomes to be valid. However, at a later stage, annealing evidently insufficient to ensure equilibration. Direct experimental in- causes an irreversible change in the overlayer properties vestigations of surface diffusion show evidence that the mo- ͑such as the diffusion parameters and adatom interaction en- bility of Dy atoms adsorbed on Mo͑112͒ in the given cover- ergy͒, so that inequality ͑1͒ reverses its sign: age range and subjected to high-temperature annealing is indeed low.9 The diffusivity equals 10Ϫ9Ϫ10Ϫ8 nm2/s at T T ϾT . ͑2͒ mob c ϭ700– 800 K. It is lower by two order of magnitude than The irreversible transitions are summarized in Fig. 4, that, e.g., for strontium films adsorbed on the same substrate which shows a wide area ͑shaded and designated as 2D at comparable coverages.10 An estimate of the time necessary glass͒ of annealing temperature in the submonolayer cover- for diffusion over the path of an inter-adatom distance, which age range at which the overlayer loses an extended order. is obtained by extrapolation of the diffusivity data to the The notations of the all ordered structures found are also temperature range optimal for the ordering of metastable ϭ 5Ϫ 3 listed in Fig. 4. Though Fig. 4 resembles a phase diagram, phases, Ta 200– 300 K, gives 10 10 s. This value is ap- that is not the case: here the overlayer state is presented at a parently distinct from the much shorter diffusion time pro- standard temperature of 100 K, contrary to a phase diagram, viding rapid equilibration of metastable phases at low tem- which shows the state at a current temperature. peratures. Hence this fact corroborates our suggestion about To prove that the effect of high-temperature annealing the dramatic drop of adatom mobility in the range ␪ on LEED patterns at low coverage is not a trivial result of ϭ0.07– 0.58 upon high-temperature annealing. The low- the removal of Dy adatoms from the surface, the work func- temperature ordering of Dy phases such as (6ϫ1) described tion changes caused by annealing were monitored. The cov- above is not the only significative fact. Precovering of the ͑ ͒ ␪Ϸ erage dependences of these changes taken after low- (Ta Mo 112 surface with a dilute Dy film ( 0.1), with its Ͻ350– 600 K) and higher-temperature annealing are pre- subsequent high-temperature annealing, strongly suppresses sented in Fig. 5. It is seen that at no value of the coverage the Sr and Li overlayer ordering, whereas without high- does high-temperature annealing return the work function to temperature annealing, the same Dy dope to the Sr or Li its value specific for the clean surface. overlayer even improves the ordering process, indicating a One of possible interpretations of the described anneal- high adatom mobility. ing irregularity might be a transition to another more stable We suggest that the origin of the hindering of the adatom phase which is characterized by too low a disordering tem- mobility by high-temperature annealing may consist in Ͻ ͑ ͒ perature, unavailable in our experiment, so that Tmob Tc adsorption-induced surface reconstruction. The Mo 112 sur- Ͻ100 K. However, the quite small variation of the work face is close-packed in the direction along the channels ͑see function ͑and thereby of the adatom dipole moment͒ due to Fig. 2͒. However, adsorption followed by high-temperature high-temperature annealing indicates that no essential weak- annealing can induce a reconstruction of its topmost layer ening of at least dipole-dipole adatom interactions happens. into some loose-packed structure, and the ordering is known Therefore, taking into account the rather wide coverage to advance very slowly on the rough surfaces.2 The rough- range in which the order is lost, we consider that even this ness of the reconstructed surface means that adatoms are interaction only can provide a sufficiently high Tc at medium mixed with surface substrate atoms, and one actually obtains coverages where the inter-adatom distances become very a kind of two-dimensional alloy. For example, such a carbon- short. induced reconstruction to a rough ‘‘checkerboard’’ structure ͑ ͒ Though some decrease of Tc should not be ruled out, is detected for Mo 112 both by LEED and STM especially at low coverages, another possibility seems to be techniques.11 Contrary to this, the mobility of Dy adatoms on more realistic: a poor order can be explained by an the original ͑non-reconstructed͒ surface Mo͑112͒ seems to be Low Temp. Phys. 27 (99–10), September–October 2001 Fedorus et al. 853 high enough to provide some ordering for all coverages even random value of the interaction energy between neighboring at temperatures as low as 100 K. Some suggestion concern- adatoms. Besides poor mobility, this is a fundamental prop- ing the mechanism of formation of a rough surface can be erty of an amorphous state. deduced from data on the work function changes ͑Fig. 5͒.Up It is worth noting that a low-coverage glassy overlayer to at least 0.5 monolayer, the curves corresponding to low- of Dy and other similar adatoms which possess an appre- temperature and high-temperature anneals are seen to run ciable magnetic moment may exhibit properties of ultrathin almost in parallel, as if 5 to 10% of the adsorbate monolayer ͑actually two-dimensional͒ amorphous magnetic materials. is no longer contributing to the work function change after the high-temperature anneal. It seems that approximately this CONCLUSION amount of dysprosium must have ‘‘reacted’’ with the surface during high-temperature annealing ͑as a random process͒,so To conclude, we have found a set of metastable ordered that this fraction is leveled down from the position of an states in two-dimensional Dy overlayers deposited onto ͑ ͒ adatom, protruding into the vacuum, to a position within the Mo 112 at low temperatures. Upon high-temperature an- two-dimensional alloy incorporated in the surface substrate nealing, they are irreversibly replaced by stable ordered layer. The result is that it makes little contribution to the structures at high coverages and by a glassy state at low and surface potential barrier. This reaction obviously happens al- medium coverages. ready with the first few percent of the monolayer, and after Support of this work by the Basic Research Foundation completion of the surface reaction ͑limited presumably by of the Ministry of Ukraine for Education and Science and by buildup of surface stress due to incorporation of Dy into the the Volkswagen-Stiftung is gratefully acknowledged. We Mo layer͒, nothing more happens. Since the surface now has thank I. Yakovkin for a critical reading of the paper. a random and rougher corrugation, Dy adsorbed on top of this two-dimensional alloy cannot order at low coverage. *E-mail: [email protected] † However, at high coverages (␪Ͼ0.58) the situation seems to Deceased **E-mail: [email protected] differ due to domination of direct lateral interactions of ada- toms, which are strong at short inter-adatom distances and switch on a more intense driving force of ordering. 1 J. M. Ziman, Models of Disorder, Cambridge Univ. Press, London ͑1979͒. Thus, the high-temperature annealing of low-coverage 2 ͑ ͒ Ͻ␪Ͻ A. G. Naumovets, Sov. Sci. Rev., Sect. A 5, 443 1984 . Dy submonolayers (0.07 0.58) results in their transition 3 O. M. Braun and V. K. Medvedev, Usp. Fiz. Nauk 157,631͑1989͓͒Sov. to a disordered state. At TϾ600 K, Dy adatoms possess a Phys. Usp. 32, 328 ͑1989͒. high mobility in this state,9 so this is actually a melted phase. 4 F. M. Gonchar, V. K. Medvedev, T. P. Smereka, and G. V. Babkin, Fiz. ͑ ͓͒ ͑ ͔͒ However, while as the temperature is lowered, the mobility Tverd. Tela 32, 1872 1990 Sov. Phys. Solid State 32,1092 1990 . 5 D. Ju¨rgens, G. Held, and H. Pfnu¨r, Surf. Sci. 303,77͑1994͒. seems to decrease dramatically before the overlayer is able to 6 A. Fedorus, D. Kolthoff, V. Koval, I. Lyuksyutov, A. G. Naumovets, and order. This is just what is typical for a vitrifying transition. H. Pfnu¨r, Phys. Rev. B 62, 2852 ͑2000͒. A peculiarity of the submonolayer vitrification is a 7 A. Fedorus, G. Godzik, V. Koval, A. Naumovets, and H. Pfnu¨r, Surf. Sci. ͑ ͒ strong influence of the substrate potential corrugation on the 460, 229 2000 . 8 I. Lyuksyutov, A. G. Naumovets, and V. Pokrovsky, Two-Dimensional behavior of an overlayer. At an arbitrary coverage, a disor- Crystals, Academic Press, New York ͑1992͒. dered film generally has an incommensurate structure. A lack 9 A. T. Loburets, A. G., Naumovets, and Yu. S. Vedula, Surf. Sci. 399,297 ͑1998͒. of overlayer mobility implies the pinning of adatoms to the 10 ͑ A. T. Loburets, N. B. Senenko, A. G. Naumovets, and Yu. S. Vedula, Phys. wells of the Peierls relief the substrate potential corrugation Low-Dimens. Semicond. Struct. 10Õ11,49͑1995͒. 8 modified by adatom lateral interaction͒. For incommensu- 11 T. McAvoy, C. Waldfried, J. Zhang, P. A. Dowben, O. Zeybek, T. Ber- rate structures, the pinning means transformation into a trams, and S. D. Barrett, Eur. Phys. J. B 14, 747 ͑2000͒. 12 phase termed chaotic.12 In this phase, the potential wells P. Bak, Rep. Prog. Phys. 45, 587 ͑1982͒. have no standard depth due to a random distribution of the This article was published in English in the original Russian journal. Repro- adatoms over the adsorption sites, and consequently, to a duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Muonium diffusion in solid CO2 D. G. Eshchenko*

Institute for Nuclear Research, Moscow 117312, Russia V. G. Storchak

Russian Research Centre ‘‘Kurchatov Institute,’’ 46 Kurchatov Sq., Moscow 123182, Russia J. H. Brewer

Canadian Institute for Advanced Research and Department of Physics and Astronomy, University of British Columbia, Vancouver, B.C., Canada V6T 1Z1 S. P. Cottrell and S. F. J. Cox**

Rutherford Appleton Laboratory, Chilton, Oxfordshire 0X11 OQX, United Kingdom E. Karlsson and R. Waeppling

Department of Physics, Uppsala University, Box 530, 751 21 Uppsala, Sweden ͑Submitted April 25, 2001͒ Fiz. Nizk. Temp. 27, 1153–1158 ͑September–October 2001͒

The quantum diffusion of interstitial muonium atoms in solid CO2 is studied in the temperature range from 5 to 200 K using the technique of muonium spin rotation and relaxation. Muonium exhibits coherent bandlike dynamics between 140 and 160 K. At low temperature Mu undergoes strong localization. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1414576͔

When positive muons are implanted into insulators they Although the concept of localization by disorder has often form muonium ͑␮ϩeϪ or Mu͒ atoms, analogous to been introduced primarily in order to describe the electronic hydrogen atoms but nearly an order of magnitude lighter.1 transport properties of solids, it may also be applied to the The dynamics of such light atoms, being intrinsically quan- quantum dynamics of heavier particles, whether charged or tum mechanical in nature, is of special interest because it neutral.4,8,9 Recent experimental results for positive muons provides critical tests of quantum diffusion theories.2–4 The (␮ϩ) in dilute metallic alloys,10 as well as for muonium basic issue in nonclassical transport is whether a wavelike or atoms (Muϭ␮ϩϩeϪ) in ‘‘dirty’’ insulators,11,12 have clearly particlelike description is appropriate, i.e., whether the tun- indicated that crystal disorder introduced by impurities dra- neling is coherent or incoherent. This depends on whether matically changes the nature of quantum diffusion for par- the interaction with the environment is such as to lead to ticles ϳ200 times heavier than the electron. In these experi- spatial localization of the wave function or to bandlike ments the environmental coupling could be varied or ͑Bloch wave͒ motion. One of the possible channels for lo- switched on and off by changing the temperature—for inter- calization of a particle is through its interaction with lattice stitial muons in aluminum metal,10 the low-energy couplings excitations ͑phonons, librons, magnons, etc.͒. In a dissipative to conduction electrons are suppressed below the supercon- environment5,6 the lattice excitations can be represented as a ducting transition, as a BCS energy gap opens at the Fermi bath of harmonic oscillators; interaction with this environ- surface, while for muonium in insulators11,12 the inelastic ment causes a crossover from coherent quantum tunneling to interactions with phonons are frozen out at low temperatures. incoherent hopping dynamics when the particle ‘‘dressed’’ Similar arguments apply for the well-known system of a di- with the lattice excitations can be effectively thought of as a lute solution of heavy 3He atoms in a 4He crystal.13,14 polaron. A common and crucial feature of all the above- At low temperatures, the environmental excitations are mentioned experiments is that the particle tunneling band- frozen out. In this case, conventional understanding suggests width ⌬ is small compared with all other relevant energy that the only possible channel for particle localization is the parameters, especially U, the characteristic shift in the parti- introduction of crystal disorder, which thus may dramatically cle’s energy level ͑relative to its value in the perfect lattice͒ change the transport properties of a solid. A well-known ex- due to crystalline disorder. For elastic strains, U is expected Ϫ3 ample is the spatial localization of electron states near the to vary as U0(a/r) with the distance r from a point defect, Fermi level in a disordered metallic system, which leads to a where U0 is the maximum shift corresponding to a site one transition into a dielectric state ͑the Anderson transition͒:7 lattice constant a away from the defect center. ͑Of course, U coherent tunneling of a particle is possible only between lev- actually has discrete values near the center; this continuous els with the same energy ͑e.g., between equivalent sites in a formulation is only meaningful when rӷa.͒ For example, ͒ ϳ 3 crystalline lattice ; in the case of strong randomness, states typical values of U0 10 K exceed muon bandwidths in with the same energy may be too spatially separated for tun- metals by 6–7 orders of magnitude; in insulators the differ- ӷ⌬ 4,15 neling to be effective. ence is not so extreme, but still U0 . Under these cir-

1063-777X/2001/27(9–10)/4/$20.00854 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 Eshchenko et al. 855 cumstances, the influence of crystalline defects extends over ics are concerned, this splitting of adjacent energy levels acts ϳ ⌬ 1/3ӷ ␰ ͑ ͒ distances R a͓U0 / ͔ a. The quantum diffusion of the as an effective disorder, creating the bias in Eq. 1 . Here interstitial particle is therefore dominated in these case by we report our study of muonium dynamics in solid CO2,in strong, long-range crystal disorder. which the anisotropic part of intermolecular interaction is so An essential characteristic of particle transport under strong that the lattice keeps its orientational order in the en- conditions of long-range crystal disorder is its inhomogene- tire solid phase. ity at low temperatures,4,16,17 where the particle dynamics Solid carbon dioxide was one of the first few solids ͑also ␶ ͒ cannot be described in terms of a single correlation time c . quartz and ice in which muonium atom formation was ob- The particle hop rate is in this case predicted4 to follow served, about 30 years ago.18 However, there was no expla- ⍀͑ ͒ nation for the mysterious muonium relaxation in CO2. The Ϫ T ␶ 1͑R͒ϰ⌬˜ 2 , ͑1͒ muonium relaxation rate was measured in a transverse mag- c 0␰2͑ ͒ϩ⍀2͑ ͒ Ϫ R T netic field of 7.2 Oe to be about 0.6ϫ106 s 1. This relaxation ⌬˜ ⍀ rate, although not high, is, however, noticeable and cannot be where 0 is the renormalized tunneling amplitude, (T)is the particle’s energy level broadening due to coupling with associated with nonuniformity of the magnetic field, inas- ␰ ϭ much as in the same experiment a few-times-lower relax- lattice excitations, and (R) a͓dU/dr͔rϭR is the difference between energy levels at two adjacent tunneling sites due to ation rate has also been measured. Nor can the muonium disorder. ͑We ignore the possibility of degenerate adjacent relaxation in CO2 be explained by its interaction with nuclear sites at the same R.͒ At low temperatures the phonon width is magnetic moments, as all of them are equal to zero in carbon reduced and dioxide. The source of muonium relaxation in solid CO2 can be ⍀͑ ͒ Ϫ T understood if muonium is anisotropic. The reason for the ␶ 1͑R͒ϰ⌬˜ 2 . ͑2͒ c 0␰2͑R͒ muonium anisotropy is that in the environment of surround- ing atoms of the medium, muonium’s electron function is not Equation ͑2͒ describes the dynamics of particles in the vicin- necessarily the spherically symmetric 1S state. The effective ity of defects, where the inhomogeneity of the crystal results ␶Ϫ1 spin Hamiltonian for static Mu in solid CO2 in an external in a spatial distribution of hop rates, c (R). Another frac- field H, assuming an anisotropic muon-electron hyperfine in- tion of particles, initially located so far from impurities or teraction, has the form19 ⌬˜ ␰ ͓ ⍀ ͔ defects that 0 exceeds (R) as well as (T) , undergoes bandlike propagation. In the limiting case of very low tem- ϭ Ϫ ␮ Ϫ ␮ ͑ ͒ H hASe•S␮ ge BSe•H g␮ ␮S␮•H, 3 peratures ͑or below a superconducting transition͒, inelastic scattering by phonons ͑or conduction electrons͒ is strongly where A is the muonium hyperfine tensor, and S, g, and ␮ are suppressed. The two fractions cannot then exchange the spins, g factors, and magnetic moments of the respective particles—they are manifest in the appearance of two distinct particles. For example, in a variety of cubic semiconductors components in the muon polarization function P(t);10–12 this muonium centers are observed with a hyperfine interaction may be taken as a signature of spatial inhomogeneity of the that is axially symmetric about any of the ͓111͔ crystal axes. particle dynamics when the disorder is described by a long- For a state of this type ͑which is the simplest case of aniso- range potential. tropic muonium͒, A is diagonal with elements Aʈ and AЌ . The foregoing discussion, like all previous theoretical The essential feature of anisotropic muonium is that at zero treatments of quantum diffusion ͑see, e.g., Refs. 4, 8, and 9͒ magnetic field the muon spin will oscillate at a frequency implicitly assumed that the diffusing particle has access to which is equal to one-half of the anisotropic part of the hy- only one site per unit cell. perfine interaction (AʈϪAЌ) divided by Planck’s constant. In In the harmonic approximation, the transport properties the general case ͑when A is a matrix of high order͒ several of a neutral particle in a simple crystalline insulator ͑e.g., a frequencies are observed.20 The amplitudes of oscillations at monoatomic or ionic crystal͒ depend only on the phonon these frequencies are functions of the angles between the modes of the lattice. For crystals composed of molecules, principal axes of A ͑which are determined by the configura- two additional contributions enter from the internal vibra- tion of muonium sites in the crystalline cell͒ and the initial tional and rotational degrees of freedom of the molecules. muon polarization. In the case of muonium diffusion this will Internal vibrations of molecules scarcely change the particle result in apparent muonium relaxation.21 The hopping sites dynamics because of their extremely high frequencies. Mo- are equivalent, but have principal axes making different lecular rotation, however, is a different matter. Two extremes angles with the muon polarization direction; thus when a Mu are possible: the molecules may rotate almost freely in the atom hops to a new site its spin begins to oscillate with crystal, or the rotational motion may be severely restricted different amplitudes for all the frequency components, result- and hence transformed into torsional excitations ͑librons͒. ing in an effective relaxation mechanism on the time scale of Since the typical rotational frequencies of molecules in crys- a mean oscillation quarter-period. To data the theory of the tals are still much higher than the particle bandwidth, in the time evolution of the anomalous muonium polarization func- first extreme the energy levels for a particle moving in dif- tion in the hopping regime is undeveloped. In quartz,21 ferent unit cells are degenerate, and therefore its dynamics anomalous muonium diffusion was described using a Kobo- remain unperturbed. In the second extreme the anisotropic Toyabe zero-field stochastic relaxation function with interaction between molecules ͑which causes orientational hopping,22 using a constant empirical ‘‘static width’’ of 4 ordering in the first place͒ changes the crystalline potential times the highest oscillation frequency. In the current paper so that this degeneracy is lifted. As far as the particle dynam- we present a way of estimating anomalous muonium hop 856 Low Temp. Phys. 27 (99–10), September–October 2001 Eshchenko et al.

FIG. 2. Temperature dependence of muonium relaxation rate in solid CO2 in ϭ ͑ ϩ ͒ weak transverse field H 2G circles: pure CO2; stars: CO2 0.1% N2O . ␮ FIG. 1. Experimental SR spectrum in pure CO2 in zero magnetic field at 40 K. latter which sets the time scale for isotropic muonium spin relaxation. In the case of anisotropic muonium, the hyperfine rates in a host with zero nuclear moments regardless of the interaction changes from site to site or the HF interaction is exact notions about the muonium hyperfine tensor and depo- modulated ‘‘by itself,’’ and the value of the muonoum hyper- larization function. fine interaction anisotropy about 2(AʈϪAЌ)/h will set the The experiments were performed on the EMu beamline time scale for the anisotropic muonium spin relaxation. To of the ISIS Pulsed Muon Facility at the Rutherford Appleton make a connection with the standard theory of muonium ͑ Ϫ5 ␦ Laboratory. Ultrahigh-purity CO2 less than 10 impurity diffusion this modulation can be presented by ‘‘effective’’ content͒ was condensed from the gas phase into a liquid and or ˜␦. In the standard theory the envelope of the muonium then frozen into a disc-shaped cell ͑24 mm in diameter and 5 precession signal can be approximated by a simple exponen- mm deep͒. Solid samples were carefully grown from the liq- tial, uid phase at typical speeds of about 5 mm/h by applying a P͑t͒ϳexp͑Ϫt/T ͒, ͑4͒ vertical temperature gradient of about 5 K across the cell. 2 Positive muons of 28 MeV/c momentum and 100% spin po- where T2 is the transverse relaxation time of the muonium ␮ϩ Ϫ1 larization were stopped in the samples, and spin rotation spin. The muonium relaxation rate T2 has a simple form in ͑SR͒ time spectra were recorded at various different tempera- two limits: if muonium ‘‘hops’’ from site to site at a rate tures and applied magnetic fields. Positrons from the muon ␶Ϫ1ӷ˜␦ ͑ ͒ c fast hopping limit , then the transverse relaxation Ϫ Ϫ decay are emitted preferentially along the direction of the rate is given by T 1Ϸ˜␦2␶ . For very slow diffusion (␶ 1 muon polarization, so that the time-differential ␮ϩSR tech- 2 c c Շ˜␦) the muonium spin relaxation takes place on a time scale nique, details of which can be found elsewhere,1 produces ␶ Ϫ1Ϸ˜␦ ͓ direct measurements of the time dependence of the muon shorter than c , and T2 . In this case P(t) is better ͔ decay asymmetry A(t), which is proportional to the muon approximated by a Gaussian relaxation function . Figure 2 presents the temperature dependences of the polarization function P(t). Ϫ1 muonium transverse relaxation rate T2 in solid CO2 and The formation of anisotropic muonium in solid CO2 was detected by observing the oscillation signal in zero magnetic CO2 plus 0.1% of N2O impurities, extracted from the spectra by fitting single-exponential relaxation functions ͑4͒. field. A typical experimental spectrum in zero magnetic field Ϫ1 at low temperature is represented in Fig. 1. Then a set of In pure CO2 at high temperatures above 160 K, T2 ␮ϩSR spectra was measured in a weak transverse magnetic decreases with decreasing temperature, or the hop rate in- field Hϭ2G. Transverse magnetic field spectra show a nice creases with decreasing temperature. This is an unambiguous anisotropic muonium precession signal. In all the tempera- manifestation of quantum diffusion which has also been rec- ͑ ͒ ture interval the envelope of this precession is approximated ognized for muonium in KCl Ref. 23 and solid nitrogen ͑ ͒ by a simple exponential function P(t)ϳexp(Ϫ␭t). Usually Ref. 24 . Such strong temperature dependences of the muo- the technique of muonium transverse relaxation is used to nium relaxation at high temperatures in solid CO2 cannot be ⍀ ͑ ͒ extract hopping rates of isotropic muonium in a host with due solely to the (T) dependence in Eq. 1 in the regime ⍀ӷ␰ nonzero nuclear moments. In this case the effective spin of dynamical destruction of the band ( ), but must Hamiltonian of muonium in a crystal in an external magnetic rather be caused by exponential renormalization of the tun- field consists of electron, muon, and nuclear Zeeman inter- neling amplitude, due to strong one-phonon scattering. actions, the muonium hyperfine ͑HF͒, and the nuclear hyper- The muonium relaxation levels off at lower temperatures ͑ ͒ ͑ ͒ below about 160 K. This plateau must represent the onset of fine NHF interactions see, e.g., Ref. 15 . Qualitatively, 4,15 modulation of the NHF interactions results in relaxation of muonium band motion, which occurs if the coherence is ⍀Ӷ⌬˜ ͑ Ͼ⌬˜ ͒ the muonium electron spin, which in turn leads to the ob- preserved, 0 but still T 0 , and the disorder is ␰Ӷ⌬˜ served muon depolarization via the muonium HF interaction. weak, 0 . As the muonium HF constant A is usually several orders of The band motion does not extend to the lowest tempera- ͑ Ϫ1 magnitude higher than the NHF interaction which for sim- tures, however: T2 begins to increase with decreasing tem- plicity we characterize by a single parameter ␦͒,itisthe perature below about 140 K. These data indicate that inter- Low Temp. Phys. 27 (99–10), September–October 2001 Eshchenko et al. 857

stitial muonium atoms undergo gradual localization in CO2 of Canada, the Engineering and Physical Sciences Research at low temperatures where ⍀Ӷ␰ and that the particle dynam- Council of the United Kingdom and the Russian Research ics follow Eq. ͑2͒. Center ‘‘Kurchatov Institute.’’ Two of us ͑V.G.S. and D.G.E.͒ ͑ Experiments with doped CO2 were performed to esti- were also supported by the INTAS Foundation through mate the muonium hopping rate from the trapping regime of grant 97-30063͒ and NATO ͑through Grant PST. ͒ muonium diffusion. Transverse relaxation data in solid CO2 CLG.977687 . with 0.1% N2O impurities are represented in Fig. 2 by stars. ͑ * When the Mu atom hops rapidly in CO2 causing low values E-mail: [email protected] Ϫ1 ͒ **And Department of Physics and Astronomy, University College London, of T2 due to dynamical ‘‘narrowing’’ ,itfindsaN2O im- ͑ Cower Street, London WC1E 6BT, UK purity in the CO2 crystal and reacts chemically probably to form the MuNNO* radical͒,25 which explains the fact that Ϫ the maximum T 1 value for Mu in CO ϩN O significantly 2 2 2 1 A. Schenck, Muon Spin Rotation: Principles and Applications in Solid exceeds that for slowly hopping Mu in pure CO2 at low State Physics, Adam Hilger, Bristol ͑1986͒;S.F.J.Cox,J.Phys.C20, temperatures, where it is determined by the anisotropy of Mu 3187 ͑1987͒;J.H.Brewer,Muon Spin Rotation/Relaxation/Resonance,in hyperfine tensor. This chemical reaction results in an expo- Encyclopedia of Applied Physics, edited by G. L. Trigg, VCH, New York nential relaxation of the Mu polarization, the rate of which is ͑1995͒. 2 A. F. Andreev and I. M. Lifshitz, Zh. E´ ksp. Teor. Fiz. 56, 2057 ͑1969͒ determined by the time required for Mu to approach the N2O ͓Sov. Phys. JETP 29, 1107 ͑1969͔͒. impurity within a distance a of about one lattice constant, 3 C. P. Flynn and A. M. Stoneham, Phys. Rev. B 1, 3966 ͑1970͒. after which the reaction occurs immediately. Then the rate of 4 Yu. M. Kagan and N. V. Prokof’ev, Quantum Tunneling Diffusion on trapping is26 Solids,inQuantum Tunneling in Condensed Media, edited by A. J. Leg- gett and Yu. M. Kagan, North-Holland ͑1992͒,p.37. ␭ϭ4␲aDc, ͑5͒ 5 A. O. Caldeira and A. J. Leggett, Phys. Rev. Lett. 46,211͑1981͒; Ann. Phys. ͑N.Y.͒ 149, 374 ͑1983͒. ϭ 2 ␶ 6 ͑ ͒ where D a /4 c is the muonium diffusion coefficient, and c U. Weiss and H. Grabert, Phys. Lett. A 108,63 1985 ; H. Grabert and U. Weiss, Phys. Rev. Lett. 54, 1065 ͑1985͒. is the concentration of impurities. From the maximum relax- 7 ͑ ͒ Ϫ1 ϳ 7 Ϫ1 ␶ ϳ␲ Ϫ1 P. W. Anderson, Phys. Rev. 109, 1492 1958 . ation rate T2 max 10 s we can estimate c max c/T2 max 8 U. Weiss, Quantum Dissipative Systems,inSeries in Modern Condensed 9 Ϫ1 ϳ3ϫ10 s . For the bandlike regime the ‘‘hop rate’’ is Matter Physics, Vol. 2, World Scientific, Singapore ͑1993͒. Ϫ 9 given by ␶ 1ϭ2&⌬˜ ͑Ref. 27͒. By substituting ␶ we T. Regelmann, L. Schimmele, and A. Seeger, Philos. Mag. B 72, 209 c 0 cmax ͑1995͒. can estimate the muonium bandwidth in CO2 to be about 10 E. Karlsson et al., Phys. Rev. B 52,6417͑1995͒. Ϫ2 10 K. This value may be compared with muonium band- 11 R. Kadono, R. F. Kiefl, W. A. MacFarlane, and S. R. Dunsiger, Phys. Rev. widths in KCl and solid nitrogen ͑0.16 and 10Ϫ2 K, B 53, 3177 ͑1996͒. ͒ 15 12 V. Storchak, J. H. Brewer, and G. D. Morris, Phys. Rev. B 53, 11300 respectively . ͑ ͒ Ϫ1 1996 . At high temperatures the clear maximum in T2 for Mu 13 M. J. Richards, J. Pope, and R. Widom, Phys. Rev. Lett. 29,708͑1972͒; ϩ in CO2 N2O marks the crossover from fast to slow Mu dif- A. R. Allen, M. G. Richards, and J. Schratter, J. Low Temp. Phys. 47,289 fusion near N O impurities. In this temperature range the ͑1982͒. 2 14 strong coupling to phonons allows Mu to overcome the de- V. N. Grigoriev, B. N. Esel’son, V. A. Mikheev, and Yu. E. Shul’man, JETP Lett. 17,25͑1973͒. fect potential and move to react with N2O. However, the 15 V. G. Storchak and N. V. Prokof’ev, Rev. Mod. Phys. 70,929͑1998͒. energy shift ␰ which the particle has to overcome is much 16 Yu. M. Kagan and N. V. Prokof’ev, Phys. Lett. A 159, 289 ͑1991͒. larger close to the defect than far from it, making the Mu hop 17 N. V. Prokof’ev, Hyperfine Interact. 85,31͑1994͒. 18 G. G. Myasishcheva et al., Zh. E´ ksp. Teor. Fiz. 53,451͑1968͓͒Sov. Phys. rate strongly dependent on the distance from the defect. JETP 26, 298 ͑1968͔͒. At low temperatures the suppression of inelastic interac- 19 B. D. Patterson, Rev. Mod. Phys. 60,69͑1988͒. tions with the lattice changes Mu diffusion drastically—Mu 20 J. H. Brewer, D. P. Spencer, D. G. Fleming, and J. A. R. Coope, Hyperfine atoms are stuck ͑or ‘‘frozen’’͒ at some distance far from N O Interact. 8, 405 ͑1981͒. 2 21 J. H. Brewer, Hyperfine Interact. 8,375͑1981͒. impurities, causing a strong reduction of the reaction rate 22 R. H. Hayano, Y. J. Uemura, J. Imazato, N. Nishida, T. Yamazaki, and ͑Mu relaxation rate͒. At very low temperature Mu atoms R. Kubo, Phys. Rev. B 20, 850 ͑1979͒. 23 simply cannot approach N2O impurities due to suppression R. F. Kieff, R. Kadono, J. H. Brewer, G. M. Luke, H. K. Yen, M. Celio, Ϫ1 and E. I. Ansaldo, Phys. Rev. Lett. 62, 792 ͑1989͒. of the inelastic interactions, and T2 becomes the same in 24 ͑ ͒ V. Storchak, J. H. Brewer, W. N. Hardy, S. R. Kreitman, and G. D. Morris, pure and doped crystals Fig. 2 . Phys. Rev. Lett. 72, 3056 ͑1994͒. In conclusion, we have presented evidence for the band- 25 J. J. Pan, D. J. Arseneau, M. Senba, M. Shelly, and D. G. Fleming, J. Phys. like motion of light interstitial muonium atoms in CO crys- Chem. A 101, 8471 ͑1997͒. 2 26 ͑ ͒ tals. T. R. Waite, Phys. Rev. 107, 463 1957 . 27 K. G. Petzinger, Phys. Rev. B 26,6530͑1982͒. This work was supported by the Canadian Institute for Advanced Research, the Natural Sciences and Engineering This article was published in English in the original Russian journal. Repro- Research Council of Canada, the National Research Council duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Search for the minimum thermal conductivity in mixed cryocrystals „CH4…1À␰Kr␰ V. A. Konstantinov,* V. G. Manzhelii, and V. P. Revyakin

B. Verkin Institute for Low Temperature Physics and Engineering of the National Academy of Sciences of Ukraine, 47 Lenin ave., Kharkov, 63103, Ukraine R. O. Pohl

Laboratory of Atomic and Solid State Physics, Cornell University, Ithaca, New York, USA ͑Submitted March 21, 2001͒ Fiz. Nizk. Temp. 27, 1159–1169 ͑September–October 2001͒

The isochoric thermal conductivity of (CH4)1Ϫ␰Kr␰ solid solutions is studied between 40 K and ϳ150 K over a wide range of concentrations ͑␰ϭ0.013, 0.032, 0.07, 0.115, 0.34, 0.71, 0.855, 0.937, and 0.97͒. A gradual transition from the thermal conductivity of a highly perfect crystal to the minimum thermal conductivity is observed as the crystal becomes increasingly more disordered. A qualitative description is given in the framework of Debye model of thermal conductivity, which takes into consideration the fact that phonon mean free path cannot decrease indefinitely. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1414577͔

INTRODUCTION The objects of this study were (CH4)1Ϫ␰Kr␰ solid solu- tions (0р␰р1). This choice was based on the following The question of what occurs when the mean free path of considerations: ͒ a phonon becomes comparable to the lattice parameter or to i The phase diagram of (CH4)1Ϫ␰Kr␰ has been studied its own wavelength is one of the most intriguing problems in quite well. It was found7 that the components form a homo- the thermal conductivity of solids.1–6 Some progress in the geneous solid solution with fcc structure above 30 K at all ␰. ͒ description of the heat transport in strongly disordered mate- ii Kr and CH4 have similar molecule/atom radii and rials has come about through the concept4–6 of the minimum parameters of the pair potential.8 The Debye temperatures ⌳ ⌰ 7,8 thermal conductivity min which is based on the picture D of Kr and CH4 are 72 and 143 K, respectively. At the where the lower limit of the conductivity is reached when the same time the masses of the Kr atom and the CH4 molecule heat is being transported through a random walk of the ther- are very different, 83.8 and 16 atomic units, respectively. The ⌬ mal energy between the neighboring atoms or molecules vi- fractional mass difference M/M is 0.81 for the CH4 impu- ⌳ brating with random phases. In this case min can be written rity in Kr and 4.23 for the Kr impurity in methane. One can as the following sum of three Debye integrals:5,6 thus expect strong impurity scattering, especially in the latter case. 1/3 2 ␪ 3 x iii͒ The peculiarities of the heat transfer in pure Kr and ␲ T i /T x e ⌳ ϭͩ ͪ 2/3 ͩ ͪ ͵ CH have been studied in detail ͑Refs. 9–12 and 13–15, k n ͚ v ͫ dxͬ. 4 min B i ⌰ ͑ xϪ ͒2 6 i i 0 e 1 respectively͒. There have been separate studies of the low- ͑ ͒ 1 temperature thermal conductivity at TϽ20 K in the mixed cryocrystals, which indicate strong point-defect 12,15 ⌳ P The summation is taken over three ͑two transverse and scattering. The thermal conductivity meas of solid CH4 ͒ ⌰ one longitudinal sound modes with the sound speeds vi , i and Kr measured at the saturated vapor pressure is shown in is the Debye cutoff frequency for each polarization expressed ⌰ ϭ ប ␲2 1/3 in degrees Kelvin, i vi( /kB)(6 n) , and n is the number density of atoms or molecules. Although no theoret- ical justification exists as yet for this picture of the heat transport, evidence for its validity has been obtained on a number of amorphous solids in which the high-temperature thermal conductivity has been found to agree with the value predicted by this model.5 Indirect evidence has also been obtained in measurements of the thermal conductivity of highly disordered crystalline solids, in which no thermal con- ductivity smaller than that predicted by this model seems to have ever been observed.6 What has been missing so far, however, has been a systematic study of the gradual transi- tion from the thermal conductivity of a highly perfect crystal to the minimum thermal conductivity as the crystal becomes ⌳ P FIG. 1. The thermal conductivities meas of solid Kr and CH4 measured at increasingly disordered. Such measurements will be pre- ⌳ saturated vapor pressures according to data of Refs. 9–15. min of solid Kr ͑ ͒ sented here. and CH4 calculated by Eq. 1 are shown at the bottom.

1063-777X/2001/27(9–10)/8/$20.00858 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 Konstantinov et al. 859

⌳ ͑ ͒ Fig. 1 together with min calculated by Eq. 1 for equilib- rium conditions. The corresponding values of the density and sound velocities at different temperatures needed for the cal- culation were taken from Refs. 8 and 16. As can be seen from Fig. 1, the thermal conductivity of Kr goes through a maximum at 12 K and then decreases, approaching its lower ⌳ limit min at premelting temperatures. The thermal conduc- tivity of CH4 goes through a maximum at 4.5 K in the ori- entationally ordered ␣ phase and decreases subsequently to the ␣-␤ transition temperature T␣Ϫ␤ϭ20.4 K. In the orien- tationally disordered ͑OD͒ ␤ phase the thermal conductivity begins to increase again, goes through another maximum at 50 K, and then decreases as far as the melting-point tempera- FIG. 3. The V-T diagram for solid Kr according to Ref. 16. The two iso- ϭ ⌳ P ⌳ 3 ture Tm 89.8 K. The ratio meas/ min is shown in Fig. 2. It chores shown correspond to the volumes 28.5 and 29.0 nm /mole. ⌳ P ⌳ is seen that meas exceeds the lower limit min by a factor of two to four in the entire region of the OD phase. For these reasons, the (CH4)1Ϫ␰Kr␰ mixed cryocrystals enable us to study the influence of atomic-size defects upon correspond to the onset of sample melting. T0 and Tm shift the thermal conductivity near its lower limit and to test the towards lower and higher temperature as the density of the validity of the concept of the minimum thermal conductivity sample increases. in the form it was reported in Refs. 5 and 6. The molar volume of (CH4)1Ϫ␰Kr␰ solid solution varies appreciably ͑about 10% at the saturated vapor pressure͒ as a function of the component concentrations.7 In this context it EXPERIMENTAL TECHNIQUE is convenient to compare the experimental results taken at To compare correctly experimental results on thermal different concentrations for those samples which reach the у⌰ ϭ conductivity with theory at T D it is necessary to perform V const condition at the same temperature T0 . There are experiments at constant density to exclude the effect of ther- certain experimental difficulties to prepare the samples with mal expansion. The simplest molecular crystals and solidi- the same temperature T0 . However, if the thermal conduc- fied inert gases are best suited, since their thermal expansion tivity is studied on several isochores at the same concentra- coefficients are much larger than those of the materials com- tion of the components, it can be recalculated to any T0 , monly used in high-pressure cells. If a high-pressure cell is using the volume dependence of thermal conductivity in the ץ ⌳ ץ ϭϪ filled with a solid sample of quite high density, on subse- form: g ( ln / ln V)T , where g is the Bridgman coeffi- ⌳ quent cooling the volume of the sample will remain practi- cient. Since the pressure at temperature T0 is zero, (T0) ⌳ ϭ cally constant while the pressure of the sample decreases. corresponds to the isobaric thermal conductivity P(P 0) The small deviations from constant volume caused by ther- as well. mal and elastic deformation of the measuring cell can easily This study was carried out on a coaxial-geometry setup, be taken into account. using the steady-state method. The measuring beryllium The V-T diagram of solid Kr plotted according to data of bronze cell ͑160 mm long, 17.6 mm inner diameter͒ was Ref. 16 is shown in Fig. 3. As an example, two isochores are engineered for a maximum pressure of 800 MPa. The inner shown, which correspond to the volumes 28.5 and 29.0 measuring cylinder was 10.2 mm in diameter. The sample 3 ϭ nm /mole. T0 is the temperature at which the V const con- was between the outer and inner cylinders. The temperature dition comes into play in the experiment. Below T0 the iso- was measured with platinum resistance thermometers, choric data convert to the data measured at saturated vapor mounted in special channels of the inner and outer cylinders pressure, and on further cooling the sample can pull away- and unaffected by high pressure. During the sample growth, ϳ from the cell walls. The temperature Tm and pressure Pm a temperature gradient 1 K/cm was maintained along the measuring cell. The pressure in the inlet capillary and the cell was produced by a special thermocompressor and was varied from 50 to 250 MPa for obtaining samples of various densities. Under identical growing conditions ͑a capillary pressure Ϸ200 MPa and a temperature gradient along the cell Ϸ1 K/cm͒ we were able to prepare denser samples of ␰Ͻ solid solutions with CH4 prevailing ( 0.3) than with Kr prevailing. This may be due to the higher compressibility of CH4 as compared to that of Kr. When the sample growth was completed, the inlet capillary was blocked at liquid hydrogen temperature and the samples were annealed at premelting temperatures for 5–6 h to remove density gradients. After completion of the experiment the samples were evaporated into a thin-wall vessel and their masses were determined by ⌳ P ⌳ weighing. The molar volumes of the samples were estimated FIG. 2. The ratio meas/ min for solid Kr and CH4. The dashed line corre- ⌳ P ⌳ ϭ sponds to meas/ min 1. from the known volume of the measuring cell and from the 860 Low Temp. Phys. 27 (99–10), September–October 2001 Konstantinov et al.

⌳ ⌳ ͑ ͒ TABLE I. Smoothed values of the isobaric P and isochoric V coefficients of thermal conductivity of the CH4 1Ϫ␰Kr␰ solid solution ␰Ͼ ͑ ͒ 3 ( 0.7) in mW/ cm•K . Molar volumes of the samples are given in nm /mole.

sample mass. The total systematic error was dominant, but The molar volumes Vm of the samples, the temperatures T0 did not exceed 4% for the thermal conductivity and 0.2% for at which the Vϭconst condition comes into play in the ex- molar volume. The purity of the Kr and CH4 gases used was periment, and the temperatures Tm of the beginning of better than 99.98%. The experimental details are described sample melting are given in Table III along with the Bridg- ץ ⌳ ץ more comprehensively in Ref. 17. ϭϪ man coefficients g ( ln / ln V)T calculated from the measured results. The isochoric thermal conductivity of pure RESULTS ͑ ͒ Kr Ref. 11 and solid solution (CH4)1Ϫ␰Kr␰ with Kr prevail- ⌳ ␰Ͼ ϭ The smoothed coefficients of isobaric P and isochoric ing ( 0.7) rescaled to T0 75 K is shown in Fig. 4 to- ⌳ 12 V thermal conductivities are presented in Tables I and II. gether with low-temperature data. Figure 5 shows the iso-

⌳ ⌳ ͑ ͒ TABLE II. Smoothed values of the isobaric P and isochoric V coefficients of thermal conductivity of the CH4 1Ϫ␰Kr␰ solid solution ␰Ͻ ͑ ͒ 3 ( 0.4) in mW/ cm•K . Molar volumes of the samples are given in cm /mole. Low Temp. Phys. 27 (99–10), September–October 2001 Konstantinov et al. 861

ϭ TABLE III. Molar volumes Vm of the samples, the temperatures T0 at which the experiment reaches the condition V const, the ץ ⌳ ץ ϭϪ temperatures Tm of the onset of sample melting, and the Bridgman coefficients g ( ln / ln V)T .

͑ ͒ choric thermal conductivity of pure CH4 Ref. 13 and of the as a whole and increases less steeply with decreasing tem- ␰Ͻ у␰у solid solution with CH4 prevailing ( 0.35), rescaled toT0 perature. At 0.3 0.7 the thermal conductivity becomes ϭ40 K, together with low-temperature data.15 Although practically temperature independent, as expected for the ⌳ у⌰ there is a temperature gap between the two sets of data, one lower limit of the thermal conductivity min at T D . The ⌳ can conclude that they agree quite well. The lower limits of measured data lie between the predicted min for pure Kr ⌳ ͑ ͒ ␰ the isochoric thermal conductivity min of pure Kr and CH4 and CH4, calculated according to Eq. 1 .As decreases ͑ ͒ are shown in Figs. 4 and 5 as broken lines. The molar vol- further CH4-rich crystal , the thermal conductivity increases umes and sound velocities needed for the calculation were again, and it approaches that of pure CH4 as the concentra- taken to correspond to the temperature T0 from Refs. 8 and tion of Kr decreases to zero. The concentration dependence 16 and were assumed to be independent of temperature. For of the thermal conductivity of the (CH4)1Ϫ␰Kr␰ solid solu- ϭ ϭ ⌳ the crystals with Kr prevailing (␰у0.97), the isochoric ther- tion at T 75 K and P 0 is shown in Fig. 6. The min of mal conductivity increases steeply as the temperature de- pure Kr and CH4 are shown as broken lines. It is seen that creases. As ␰ decreases, the thermal conductivity decreases

FIG. 4. Smoothed values of the isochoric thermal conductivity of pure Kr FIG. 5. Smoothed values of the isochoric thermal conductivity of pure CH4 ͑ ͒ ϭ ᭿ ͑ ͒ ϭ ᭿ Ref. 11 and Kr with a CH4 admixture, rescaled to T0 75 K( ), together Ref. 13 and CH4 with a Kr admixture, rescaled to T0 40 K( ), together 12 ⌳ 15 ⌳ with low-temperature data. The broken lines are the isochoric min of pure with low-temperature data. The broken lines are the isochoric min of pure Kr and CH4. Kr and CH4. 862 Low Temp. Phys. 27 (99–10), September–October 2001 Konstantinov et al.

FIG. 6. The concentration dependence of the thermal conductivity of FIG. 7. The isochoric thermal resistivity of pure Kr and Kr with a CH ϭ ϭ 4 (CH4)1Ϫ␰Kr␰ solid solutions at T 75 K and P 0. The horizontal lines are admixture, rescaled to T ϭ75 K. The solid and dashed lines correspond to ⌳ 0 min of pure Kr and CH4 under the same conditions. the best fit within the Debye model of the thermal conductivity according to i͒ and ii͒, respectively. thermal conductivity changes rapidly at ␰Ͻ0.2 and ␰Ͼ0.8 and is practically concentration independent at 0.2Ͻ␰Ͻ0.8. scattering.13,14 However, it decreases rapidly when the tem- perature increases and it practically disappears above 90 K, whereas the phonon-phonon scattering continues to increase. DISCUSSION The competition of these two ͑phonon-phonon and phonon- In the solidified inert gases Ar, Kr, and Xe the phonon– rotational͒ mechanisms of scattering in the OD phase of solid phonon interaction is the main mechanism determining the CH4 results in another high-temperature maximum of the value and the temperature dependence of the thermal con- thermal conductivity, in addition to the low-temperature one ductivity ⌳(T) at the Debye temperature and higher ͑He and at 4.5 K. The models describing the phonon scattering on ⌰ ͑ Ne melt at temperatures much below D Ref. 16, Chapter fluctuations of the orientational short-range order in solid 11͒͒. If the scattering is not too strong and the model of methane were suggested in Refs. 14 and 22. It has already 1–4,18 ⌳ ⌳ elastic waves is appropriate, the theory predicts that the been mentioned above that meas exceeds min by a factor of ⌳ ϰ isochoric thermal conductivity should vary as V 1/T at T two to four in the OD phase of methane. The additional у⌰ ͑ D . However, experiments at constant volume revealed scattering mechanism due to Kr impurity which is supposed appreciable deviations from this dependence at the highest to be very strong because of the big mass difference͒ de- temperatures, with the conductivity varying slower than creases the thermal conductivity with increasing ␰ to values 11,16,19 3,19 ⌳ 1/T. Such behavior implies an approach of the ther- which are between the predicted min for pure Kr and CH4, ⌳ ͑ ͒ mal conductivity toward its lower limit min . The proximity calculated according to Eq. 1 . ⌳ of the thermal conductivity of solidified inert gases to min at Defects of atomic dimensions weaken the temperature premelting temperatures was first pointed out by Slack.4 dependence of the lattice thermal conductivity caused by A characteristic peculiarity of molecular crystals is the U-processes.3,23 The theoretical consideration of point defect у⌰ presence of both translational and rotational motions of mol- scattering in the framework of the Debye model at T D is ecules in their lattice sites. For normal ͑orientationally or- made with the assumption that the perturbation theory is dered͒ molecular solids the intermolecular modes generally valid at all frequencies and temperatures.18,23,24 Point defects have a mixed translational–librational character and cannot at small concentrations (␰Ͻ0.05) have to lead to an addi- be considered as independent. It is assumed, however, that tional contribution to the thermal resistivity Wϭ1/⌳ of the separation can be made in orientationally ordered phases at crystal, which is independent of temperature.24 The slowest low temperatures, and then collective translational vibrations variation in the limit of very strong point-defect scattering is ͑phonons͒ and rotational vibrations ͑librons͒ can be treated ⌳ϰ1/T1/2 ͑Ref. 25͒. It has been already noted above that the separately.20 The contribution of librons to the heat transfer thermal conductivity of Kr appears to approach its lower ⌳ ͑ ͒ ⌳ is assumed to be relatively small because the bands associ- limit min at premelting temperatures see Fig. 1 .If min ated with the angular motion of the molecules are narrow. At actually is the lower limit to the thermal conductivity of a the same time, the role of librons in scattering processes is lattice, the point-defect scattering should be different from important.20 In orientationally disordered phases, however, that discussed above. Intuitively, one would expect that the the absence of orientational long-range order means that contribution of small concentrations of point defects to the well-defined, librational modes cannot propagate through the thermal resistivity should decrease with increasing tempera- crystal: they are always rapidly damped out.21 The additional ture. In the case of very strong point-defect scattering ⌳ may ͑ ͒ ⌳ phonon scattering beyond the phonon-phonon may origi- coincide with min in a wide temperature range. To check nate in the OD phases of molecular crystals due to the short- this assumption the thermal resistivity of pure Kr and range fluctuations of the orientational order in the vicinity of (CH4)1Ϫ␰Kr␰ solid solutions with Kr prevailing is shown in Ϸ the phase transition. For example, in solid CH4 this addi- Fig. 7 (T0 75 K). It is seen clearly that the additional con- ␣ ␤ tional scattering above the - transition temperature T␣Ϫ␤ tribution of CH4 impurities to the total thermal resistivity ϭ20.4 K is of the same magnitude as the phonon-phonon decreases with increasing temperature at all ␰Ͼ0.7. At 0.85 Low Temp. Phys. 27 (99–10), September–October 2001 Konstantinov et al. 863

у␰у0.7 the thermal conductivity lies between the predicted the case of U-processes only.3 Let us extend this consider- ⌳ ͑ ͒ min for pure Kr and CH4, calculated according to Eq. 1 ation to the case of combined U-processes and Rayleigh scat- ͑see Fig. 4͒. tering. A quantitative description of the thermal conductivity of i͒ Let us assume that the lower limit of the phonon mean ␰Ͼ (CH4)1Ϫ␰Kr␰ solid solutions with Kr prevailing ( 0.7) is free path is of the order of one-half of the phonon wave- made by taking into consideration only U-processes and length. In this case l(x) is given by Eq. ͑3͒ when l(x) Rayleigh scattering. Phonon scattering on excitations of the у␣␭/2 or otherwise by rotational motion of the methane molecules, which play an l͑x͒ϭ␣␭/2, ͑8͒ important role at low temperature,12 is expected to be weak at high temperatures. It was shown earlier that molecules where ␣ is some numerical coefficient of the order of unity ϳ ␭ undergo a almost free rotation in pure CH4 at T 90 K and which is assumed to be independent of T and . The cross- higher,13 and that this kind of rotation does not lead to addi- over from behavior described by Eq. ͑3͒ to Eq. ͑8͒ occurs at the phonon wavelength ␭ , which is determined from tional phonon scattering. For matrix-isolated molecules in * solidified inert gases the onset of almost free rotation takes CT B Ϫ1 ␭ place at lower temperatures.8 ͩ ϩ ͪ ϭ␣ * . ͑9͒ ␭2 ␭4 2 The thermal conductivity can be described using the De- * * bye model as This corresponds to the effective crossover temperature ⌰ ϭ ␭ ͑ ⌰ Ͻ⌰ ⌰* 3 ⌰ 4 x hv/kB . It is supposed that D ; otherwise T D /T x e ϭ⌰ ͒ * * * ⌳͑ ͒ϭ ͩ ͪ ͵ ͑ ͒ ͑ ͒ D . The thermal conductivity integral is broken into two T 3nkBv ⌰ l x ͑ xϪ ͒2 dx, 2 D 0 e 1 parts: where ⌰ ϭv(ប/k )(6␲2n)1/3 is the Debye temperature, v T 3 ⌰ /T x4ex D B ⌳͑T͒ϭ3nk vͩ ͪ ͫ ͵ * l͑x͒ dx B ⌰ xϪ ͒2 is the polarization-averaged sound velocity, and l(x) is the D 0 ͑e 1 у⌰ phonon mean free path. At T D the boundary scattering is ⌰ 4 x negligible and the phonon mean free path is determined by D /T ␭ x e ϩ ͵ ␣ dxͬ. ͑10͒ 18 x 2 the U-processes and Rayleigh scattering: ⌰ /T 2 ͑e Ϫ1͒ * ͑ ͒ϭ͑ Ϫ1 ϩ Ϫ1 ͒Ϫ1 ͑ ͒ у⌰ l x lRayl lUmkl , 3 In the high-temperature limit T D , when only the ⌳ ϭ 2ϩ⌳Ј ͑ U-processes are in effect, (T) A1 /T min Refs. 3 and where xϭvh/k T␭, and ␭ is the phonon wavelength. ͒ B 27 , where A1 is some temperature-independent numerical Frequently used expressions for the phonon mean free ⌳Ј coefficient, and min is path resulting from the individual scattering mechanisms are 1/3 2 ⌰ 3 x ␲ T D /T x e ⌳Ј ϭ ␣ͩ ͪ 2/3 ͩ ͪ ͵ 3 n kBv dx. Ϫ CT min 6 ⌰ ͑exϪ1͒2 l 1 ϭ , ͑4͒ D 0 Umkl ␭2 ͑11͒ ⌳Ј where This expression for min is identical to that predicted by Eq. ͑1͒ if the polarization-averaged speed of sound is used and it 12␲3 ␥2k is assumed that ␣ϭ1. In the case of strong phonon scattering ϭ B ͑ ͒ C 2 5 caused by point defects ͑when ⌰ Ӷ⌰ ͒, the lower limit on &n1/3 mv * D the thermal conductivity is accurately found from Eq. ͑11͒. and ii͒ Let us assume that the lower limit of the phonon mean free path is of the order of the interatomic distance. In ␲3⍀ Ϫ1 4 0T this case the phonon mean free path l(x) is given by Eq. ͑3͒ l ϭ , ͑6͒ Ϫ Ray1 ␭4 when l(x)у␣n 1/3 or otherwise by ϭ␣ Ϫ1/3 ͑ ͒ where l͑x͒ n . 12 The crossover from behavior described by Eq. ͑3͒ to Eq. ⌬M ⌬⍀ 2 ⌫ϭ␰͑ Ϫ␰͒ ϩ ␥ 0 ͑ ͒ ͑12͒ occurs at the phonon wavelength ␭ , which is deter- 1 ͫ 2 ⍀ ͬ . 7 * M 0 mined from Here ␥ is the Gru¨neisen parameter, taken as 2.5 for Kr and CT B Ϫ1 ͑ ͒ ͩ ϩ ͪ ϭ␣nϪ1/3. ͑13͒ CH4 Refs. 8 and 16 , m is the atomic/molecular mass, and M ␭2 ␭4 ⍀ and 0 are the average atomic/molecular weight and vol- * * ͑ ⍀ The effective crossing temperature is ⌰ ϭhv/k ␭ . ͑It is ume, respectively, of the mixed crystals for pure CH4 0 * B * ϭ ϫ Ϫ23 3 ϭ ⍀ ϭ supposed, just as in the previous case, that ⌰ Ͻ⌰ ; other- 5.18 10 nm , M 16, and for pure Kr 0 4.73 * D ϫ Ϫ23 3 ϭ ͒ ⌬ ⌬⍀ wise ⌰ ϭ⌰ .͒ The thermal conductivity is again a sum of 10 nm , M 83.8 , and M and 0 are the atom/ * D molecule weight and volume difference from their average two integrals: values, respectively. T 3 ⌰ /T x4ex Expression ͑3͒ cannot apply when l(x) becomes less ⌳͑T͒ϭ3nk vͩ ͪ ͫ ͵ * l͑x͒ dx B ⌰ ͑exϪ1͒2 than the interatomic spacing or shorter than the phonon D 0 wavelength. In that case a lower limit of the mean free path ⌰ 4 x D /T x e ϩ␣nϪ1/3 ͵ dxͬ. ͑14͒ is reached. This problem has been discussed previously for x 2 ⌰ /T ͑e Ϫ1͒ * 864 Low Temp. Phys. 27 (99–10), September–October 2001 Konstantinov et al.

͑ ͒ ͒ ͒ TABLE IV. Parameters of the Debye model fits for CH4 1Ϫ␰Kr␰ solid solutions for cases i and ii ,

i͒ ii͒

␰ ϫ Ϫ22 Ϫ3 ⌫ ␣ ␣ n 10 ,cm v,km/s calc C,cm/K C, cm/K 1 2.11 0.86 0 1.07ϫ10Ϫ9 1.29 0.95ϫ10Ϫ9 1.16 0.97 2.10 0.88 0.055 1.17ϫ10Ϫ9 1.22 1.07ϫ10Ϫ9 1.24 0.937 2.09 0.92 0.12 2.03ϫ10Ϫ9 1.41 1.63ϫ10Ϫ9 1.26 0.855 2.06 0.97 0.26 3.5ϫ10Ϫ9 1.3 3.8ϫ10Ϫ9 1.20 0.71 2.04 1.07 0.54 6.2ϫ10Ϫ9 1.05 4.5ϫ10Ϫ9 0.97

у⌰ In the high-temperature limit T D , when only the reduce the number of variable parameters, the coefficient B ⌳ ϭ 3/2ϩ⌳Љ ͑ U-processes are in effect, (T) A2 /T min Refs. 3 describing the Rayleigh scattering was calculated in accor- ͒ ⌳Љ ͑ ͒ ͑ ͒ and 27 , where A2 is some numerical coefficient, and min is dance with Eqs. 6 and 7 . The Debye model parameters ⌫ ␣ 3 ⌰ 4 x used for fitting (n,v, ) and the C and values obtained by T D /T x e ⌳Љ ϭ ␣ 2/3 ͩ ͪ ͵ ͑ ͒ fit for variants i͒ and ii͒ are shown in Table IV. When Eqs. min 3 n kBv ⌰ ͑ xϪ ͒2 dx. 15 D 0 e 1 ͑16͒ and ͑17͒ are used for the phonon mean free path, we ␣Ϸ In the case of strong phonon scattering by point defects obtain 0.3– 0.4, while elementary considerations lead to ͑ ⌰ Ӷ⌰ ͒ ␣у1. The coefficient C determining the intensity of the when D , the lower limit on the thermal conductivity * ͑ ͒ U-processes is in good agreement with low-temperature is accurately found from Eq. 15 . It can be shown easily that Ϫ ⌳Љ ⌳Ј ϭ у⌰ ␣ data,12 where it was evaluated as 1.5ϫ10 9 cm/K. the ratio min/ min 0.83 at T D , for the same . iii͒ We also consider the case when a term ␣␭/2 or The fitting results are shown in Fig. 7. It is seen that both ␣nϪ1/3 has simply been included in Eq. ͑3͒ for the phonon the variants describe well the behavior of the isochoric ther- mean free path to eliminate the nonphysical case when l(x) mal conductivity of the solid solution, except for the concen- ␰ϭ ͒ becomes smaller than one-half the wavelength or than the tration 0.71. In the case of pure Kr variant ii ensures a interatomic distance: somewhat better fit, as was discussed earlier for the isochore ϭ 27 3 with T0 0K. Roufosse and Klemens preferred assump- ͑ ͒ϭ͑ Ϫ1 ϩ Ϫ1 ͒Ϫ1ϩ␣␭ ͑ ͒ l x lUmkl lRayl /2, 16 tion ii͒ to i͒ for the following reason. Assumption i͒ means Ϫ Ϫ Ϫ that the vibrational excitations can be described as waves. l͑x͒ϭ͑l 1 ϩl1 ͒ 1ϩ␣n 1/3, ͑17͒ Umkl Rayl However, when the wave picture is questioned, the alterna- and then Eq. ͑2͒ is used to calculate the thermal tive description assumes localized excitations that hop from 26 у⌰ conductivity. In the high-temperature limit T D , this site to site. The mean free path for those excitations is given ⌳ϭ ϩ⌳Ј assumption leads to the expressions: A3 /T min and by the interatomic distance. ⌳ϭ ϩ⌳Љ ͑ ͒ A4 /T min Ref. 27 , where A3 and A4 are some nu- This point of view may hold for amorphous materials as merical coefficients. well, in which the heat transfer by low-frequency phonons is The computer fit using Eqs. ͑10͒, ͑14͒, ͑16͒, and ͑17͒ negligible at quite high ͑above 50 K͒ temperatures.28 We was performed for the isochoric thermal conductivity of pure should remember that in our case the model assumes a ͑␰ϭ Kr and (CH4)1Ϫ␰Kr␰ solid solutions 0.97, 0.937, 0.855, gradual change from pure extended-phonon thermal conduc- ͒ ϭ 0.71 for the isochores with T0 75 K. The density was es- tivity ͑a perfect crystal and low temperatures͒ to a heat trans- timated by interpolation of experimental data. The sound ve- fer by localized excitations ͑a solid solution in the middle of locity was calculated assuming that it changes linearly with the composition range and at high temperatures͒. We believe ͑ ͒ ͑ ͒ the concentration from Kr Ref. 16 to CH4 Ref. 8 .To it is more reasonable to consider the vibrations localized in the ␭/2 regions as the limiting case of the phonon picture ͑see Refs. 5 and 6͒.

FIG. 8. The temperature dependence ␻ /␻ , where ␻ ϭ2␲v/␭ is the * D * * frequency at which the crossover from the wave picture to the picture of FIG. 9. The relative extended-phonon contribution to the thermal conduc- localized excitations occurs. tivity of pure Kr and the (CH4)1Ϫ␰Kr␰ solid solution. Low Temp. Phys. 27 (99–10), September–October 2001 Konstantinov et al. 865

proper description of the isochoric thermal conductivity of the (CH4)1Ϫ␰Kr␰ solid solution in the middle of composition range 0.2Ͻ␰Ͻ0.8.

The authors would like to acknowledge B. Ya. Gorodilov and A. I. Krivchikov for helpful discussions. This work was supported by National Research Council of USA under Twinning Program 1999–2000.

*E-mail: [email protected]

1 P. B. Allen and J. L. Feldman, Phys. Rev. B 48, 12581 ͑1993͒. 2 P. Sheng, M. Zhou, and Z. Zhang, Phys. Rev. Lett. 72,234͑1994͒. 3 M. C. Roufosse and P. G. Klemens, J. Geophys. Res. 79,703͑1974͒. FIG. 10. The concentration dependence of the Bridgman coefficient gϭ 4 G. A. Slack, in Solid State Physics, Vol. 34, edited by F. Seitz and ͒ ͑ ϭ ץ ⌳ ץ Ϫ ( ln / ln V)T of the solid solution (CH4)1Ϫ␰Kr␰ at T 90 K. D. Turnbull, Academic Press, New York 1979 . 5 D. G. Cahill and R. O. Pohl, Solid State Commun. 70,927͑1989͒. 6 D. G. Cahill, S. K. Watson, and R. O. Pohl, Phys. Rev. B 46, 6131 ͑1992͒. 7 Figure 8 shows the temperature dependence of ␻ /␻ , V. G. Manzhelii, A. I. Prokhvatilov, I. Ya. Minchina, and L. D. Yantsevich, * D Handbook of Binary Solutions of Cryocrystals, Begell House Inc., N. Y., where ␻ ϭ2␲v/␭ is the frequency at which the crossover * * Wallingford, UK ͑1996͒. from the wave picture to the picture of localized excitations 8 V. G. Manzhelii, A. I. Prokhvatilov, V. G. Gavrilko, and A. P. Isakina, occurs. Equation ͑10͒ actually describes the sum of contribu- Structure and Thermodynamic Properties of Cryocrystals, Begell House ͑ ͒ tions to the thermal conductivity resulting from extended Inc., N.Y., Wallingford, UK 1999 . 9 I. N. Krupskii and V. G. Manzhelii, Sov. Phys. JETP 55,2075͑1968͒. phonons and localized excitations. Figure 9 presents the rela- 10 P. Korpium, J. Moser, F. J. Piezinger, and E. Lusher, in Phonon Scattering tive extended-phonon contribution to the thermal conductiv- in Solids, edited by L. J. Challis, V. W. Rampton, and A. F. G. Wyatt, Plenum Press, New York, and London ͑1975͒,p.377. ity of pure Kr and the (CH4)1Ϫ␰Kr␰ solid solution. It is seen 11 A. I. Bondarenko, V. G. Manzhelii, V. A. Popov, M. A. Strzhemechnyi, that in Kr the thermal conductivity is of pure extended- and V. G. Gavrilko, Fiz. Nizk. Temp. 86, 1215 ͑1982͓͒Sov. J. Low Temp. phonon character up to about 90 K. As the CH4 concentration Phys. 8, 617 ͑1982͔͒. grows, progressively more heat is transferred by the local- 12 V. V. Dudkin, B. Ya. Gorodilov, A. I. Krivchicov, and V. G. Manzhelii, Fiz. ͑ ͒ Nizk. Temp. 26, 1023 ͑2000͓͒Low Temp. Phys. 26, 762 ͑2000͔͒. ized excitations, but even at the highest concentration 29% 13 ͑ ͒ V. A. Konstantinov, V. G. Manzhelii, V. P. Revyakin, and S. A. Smirnov, and the highest temperatures 150 K an appreciable part of Physica B 262, 421 ͑1999͒. the heat ͑about 10%͒ is transferred by the low-frequency 14 V. G. Manzhelii and I. N. Krupskii, Fiz. Tverd. Tela ͑Leningrad͒ 10,284 extended phonons. The values of ␣ vary from 1 to 1.4. This ͑1968͓͒Sov. Phys. Solid State 10,221͑1968͔͒. 15 B. Ya. Gorodilov, V. V. Sumarokov, P. Stachowiak, and A. Jezovski, Phys. supports the interpretation of the vibrations being localized ͑ ͒ ␭ Rev. B 58, 3089 1998 . in /2 regions as the limiting case of the phonon picture and 16 Rare Gas Solids, V. II, edited by M. L. Klein and J. A. Venables ͑Eds.͒, the validity of Eq. ͑1͒ for prediction of the lower limit to the Academic Press, London-New York ͑1977͒. thermal conductivity of the crystalline lattice. 17 V. A. Konstantinov, S. A. Smirnov, and V. P. Revyakin, Prib. Tekh. Exp. 42,145͑1999͓͒Instrum. Exp. Tech. 42,133͑1999͔͒. The concentration dependence of the Bridgman coeffi- 18 R. Berman, Thermal Conduction in Solids, Oxford, Clarendon Press ץ ␭ ץ ϭϪ cient g ( ln / ln V)T is shown in Fig. 10. It is seen that ͑1976͒. the value of g is highest for pure components and decreases 19 V. A. Konstantinov, V. G. Manzhelii, M. A. Strzhemechnyi, and S. A. strongly for solid solutions. The same behavior was also ob- Smirnov, Fiz. Nizk. Temp. 14,90͑1988͓͒Sov. J. Low Temp. Phys. 14,48 ͑ ͔͒ served for mixed alkali-halide crystals,29 though the decrease 1988 . 20 V. G. Manzhelii, V. B. Kokshenev, L. A. Koloskova, and I. N. Krupskii, in g was not as large. Small Bridgman coefficients (g Fiz. Nizk. Temp. 1,1302͑1975͓͒Sov. J. Low Temp. Phys. 1,624͑1975͔͒. Ϸ3 – 4) are typical for amorphous materials and strongly 21 B. M. Powell and G. Dolling, Can. J. Chem. 66, 897 ͑1998͒. disordered crystals.30 22 H. Yasuda, J. Low Temp. Phys. 31,223͑1978͒. 23 P. G. Klemens, Phys. Rev. 119, 507 ͑1960͒. 24 V. S. Oscotskii and L. A. Smirnov, Defects in Crystals and Heat Conduc- CONCLUSION tion ͓in Russian͔, Nauka, Leningrad ͑1972͒. 25 L. A. Turk and P. G. Klemens, Phys. Rev. B 9, 4422 ͑1974͒. A gradual transition from the thermal conductivity of a 26 J. R. Olson, R. O. Pohl, J. W. Vandersande, A. Zoltan, T. R. Anthony, and highly perfect crystal to the minimum thermal conductivity W. F. Banholzer, Phys. Rev. B 47, 14850 ͑1993͒. 27 V. A. Konstantinov, J. Low Temp. Phys. 122, 459 ͑2001͒. ⌳ is observed at Tу⌰ in the (CH ) Ϫ␰Kr␰ solid solu- min D 4 1 28 M. S. Love and A. S. Anderson, Phys. Rev. B 42, 1845 ͑1990͒. tions as the crystal becomes increasingly disordered. A quali- 29 D. Gerlich, J. Phys. C 20, 5479 ͑1987͒. tative description is given in the framework of Debye model 30 R. G. Ross, P. A. Andersson, B. Sundqvist, and G. Backstrom, Rep. Prog. of thermal conductivity, which takes into consideration the Phys. 47, 1347 ͑1984͒. fact that phonon mean free path cannot decrease indefinitely. This article was published in English in the original Russian journal. Repro- The concept of the minimum thermal conductivity3,4 gives a duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Vibrational dynamics of solid molecular nitrogen to megabar pressures A. F. Goncharov,* E. Gregoryanz, H.-K. Mao, and R. J. Hemley

Geophysical Laboratory and Center for High Pressure Research, Carnegie Institution of Washington, 5251 Broad Branch Road, N. W., Washington, D. C. 20015 ͑Submitted July 2, 2001͒ Fiz. Nizk. Temp. 27, 1170–1174 ͑September–October 2001͒ We report the results of Raman and synchrotron infrared absorption measurements of several molecular phases of solid nitrogen to pressures above 100 GPa ͑300 K͒. Low- temperature vibrational spectra to 45 GPa are also presented. The changes in Raman and infrared spectra at 60 GPa and 300 K are interpreted as arising from the ␧→␨ transition reported at ␯ low temperature. The character of splitting of the Raman vibron 2 observed at 25 GPa and low temperatures differs from that previously reported, a difference that we ascribe to different experimental procedures employed and metastability of the low-temperature phase. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1414578͔

1. INTRODUCTION cially anisotropic terms͒ to obtain adequate results. At 20–25 GPa and 10 K a transformation has been found With the recent progress in diamond-anvil cell tech- by Raman spectroscopy21 and confirmed by later Raman and niques ͑see Ref. 1͒, a growing number of examples of orien- IR measurements.33 No x-ray data are currently available for tational ordering and subsequent transformations to frame- this phase ͑called ␨͒. According to vibrational spectroscopy work structures in pressurized molecular solids have been data, its structure has a strong similarity to phases ␧ and ␦.It documented.2–6 For transformation to framework structure, has been inferred to have R3c symmetry,21 but more recent information about the type and character of the ordering is observations do not support this interpretation.33 The ␧–␨ important, since it affects the energy barriers associated with phase boundary has been extended recently to 180 K and 40 the transition. Solid nitrogen is an important system for such GPa.33 At low temperatures, a significant region of metasta- studies because of stability and simple electronic structure of bility is reported,33 making it difficult to clarify the mecha- the isolated molecule. Moreover, solid nitrogen has been nism of the transformation. well studied theoretically, and accurate experimental data At room temperature, the vibrational spectroscopy data provide an important test of condensed matter theory.7–12 are still contradictory concerning the number and nature of Solid nitrogen has a complex phase diagram with a the transformations above 20 GPa. A sequence of new phases wealth of molecular phases that differ in the types of orien- has been reported on the basis of several branchings of the 22 27 tational ordering and crystal structures formed.5,13–33 The or- Raman vibron modes, including one just above 20 GPa. dering of the low-pressure, low-temperature ␣ and ␥ phases In contrast, x-ray studies indicate the stability of ␧ phase to 26,30 is controlled by quadrupole–quadrupole interactions, 50 GPa. A change of x-ray diffraction pattern was ob- 25 whereas at higher pressures a class of molecular structures served above 60 GPa, but interpretation requires additional ͑␦ ␦ ␧ ␨͒ measurements. The most-recent Raman and IR measure- , loc , , stabilized by additional anisotropic intermo- lecular interactions is found.34 The ␦ phase is proposed to ments to 42 GPa show a clear correspondence between the have a disordered cubic structure ͑space group Pm3n͒20 with number of observed lattice and vibron modes and group- 33 two different types of molecules exhibiting spherelike and theoretical predictions for the ␧ phase. Here we report the disklike disorder and giving rise to two classes of vibron results of a combined Raman and synchrotron IR high- ␯ ␯ pressure study of N to 40 GPa at low temperatures and bands 1 and 2 . With increasing pressure and/or decreasing 2 temperature, nitrogen molecules exhibit orientational order beyond 100 GPa at room temperature. ͑ ␦ through a sequence of phase transformations to loc and then ␧͒ as determined by vibrational spectroscopy and x-ray 2. EXPERIMENTAL 27–31 ␧ diffraction. The structure of the phase is rhombohedral Samples of nitrogen were loaded in diamond anvil cells ͑space group R¯3c͒, which can be viewed as a distortion of at 0.2 GPa using a standard gas-loading technique. The low- 23 cubic. temperature experiment was performed at the NSLS ͑beam- Theoretical calculations have been contradictory in their line U2A͒. The description of our experimental setup has 10–12 predictions of stable molecular ordered phases. In par- been published elsewhere.37 The typical spectral resolution 10 11 ticular, Monte Carlo ͑MC͒ and molecular dynamics ͑MD͒ was about 4 nmϪ1 in both IR, and Raman measurements. A simulations based on a proposed intermolecular potential fa- diamond cell containing the sample at 8 GPa was cooled to 12 vor tetragonal structures, while ab initio calculations pre- 15 K in a continuous-flow He cryostat. The cryostat is dict the stability of the R¯3c phase, in agreement with experi- equipped with windows that are interchangeable at low tem- ment. Recent MC calculations35,36 have pointed out the perature, which allows in situ IR and Raman/ruby fluores- importance of using accurate atom–atom potentials ͑espe- cence measurements without changing the P – T conditions

1063-777X/2001/27(9–10)/4/$20.00866 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 Goncharov et al. 867

FIG. 1. Raman spectra of nitrogen through the low-temperature transition to FIG. 2. Infrared absorption spectra of nitrogen at 15 K as a function of the ␨ phase. The spectra are shifted in vertically for clarity. The spectrum at pressure. The spectra are shifted in vertically for clarity. 24.8 GPa ͑points͒ is shown along with the fitting to multiple components ͑Voigt profiles͒. ␯ dependence of the frequency of the Raman 2 multiplet will on the sample.38 Pressure can be controlled from the outside be discussed later. The pressure dependences of the Raman by a mechanical change in the load on the diamond cell lever and infrared frequencies are compared in the inset of Fig. 3. arm through vacuum feedthroughs. In the room-temperature We find that Raman and infrared frequencies do not coincide ͑ ͒ experiments we used diamond anvil cells with external heat- cf. Ref. 33 . Unlike the Raman data, our infrared frequen- ing capabilities. This allows us to release the internal stresses cies are in good agreement with those reported in Ref. 33 in ͑ that build up during a loading of the cell by heating the the pressure range overlapped by the two studies to 25.2 ͒ sample to 200–400 °C. GPa . Raman spectra measured close to room temperature are presented in Fig. 4. At 60 GPa Raman spectra already con- 3. RESULTS tain more vibron bands ͑four͒ than is allowed by group The sequence of Raman spectra measured as a function ¯ ␧ ͑ ͒ 21 theory for the R3c symmetry of -N2 three . Surprisingly, of pressure at low temperatures is shown in Fig. 1. At 17.5 at this pressure we did not observe the low-frequency shoul- GPa we observe two Raman peaks in the vibron region— der of the most intense ␯ Raman band, as reported in Ref. ␯ ͑ ͒ ␯ 2c slightly broadened 2 lower frequency and 1 . This is in 32. Increasing the pressure through the 60 GPa range gives excellent accord with the experiments for the ␧ phase ͑see also Refs. 21 and 33͒. Low-frequency lattice modes ͑not shown͒ are also in agreement with previous studies21,33 for ␧ ␯ -N2. Increasing pressure splits the 2 peak, so three com- ponents can be seen. This splitting becomes obvious at the highest pressure ͑44 GPa͒, while at 24–38 GPa a peak-fitting is required to reveal the two components ͑e.g., the spectrum ͒ ␯ at 24.8 GPa in Fig. 1 . Only a slight broadening of the 1 peak is observed as the pressure is increased. Lattice modes also change dramatically at 18–25 GPa, which is beyond the scope of the present paper. The infrared spectra at 17.5 GPa ͑Fig. 2͒ show a very ␯ weak absorption in the spectral range of the 2 vibron. The absorbance increases substantially at 18–25 GPa, so a dou- blet of IR vibrons is clearly visible at higher pressure. As in the case of the Raman bands, a moderate broadening of the IR peaks is observed as the pressure is increased. The pressure dependence of the observed Raman fre- quencies is shown in Fig. 3. Corresponding data from Ref. FIG. 3. Raman vibrational frequencies as a function of pressure at 15 K. 21 ␯ 33 are also shown for comparison. Earlier results for the 2 Filled circles are the Raman data from this work. The solid thick solid lines multiplet are very close to those reported in Ref. 33 and are are the guides to the eye. The dashed lines are the Raman data from Ref. 33. not shown for clarity. For the ␯ band, our low-temperature The inset shows the comparison between Raman and infrared frequencies. 1 Filled squares are the infrared data from this work. The solid thick solid data are very close to those measured at room temperature lines are the guides to the eye. Thin solid lines are the Raman data from this ͑see also Ref. 32͒. The substantial difference in the pressure work. The dashed lines are the infrared data from Ref. 33. 868 Low Temp. Phys. 27 (99–10), September–October 2001 Goncharov et al.

nonmolecular phase. This behavior will be examined in de- tail elsewhere.

4. DISCUSSION The splitting of the vibron bands and change in the lattice-mode spectrum observed at low temperature indicate a phase transformation related to orientational ordering of the nitrogen molecules. The phase diagram of Ref. 33 sug- gests that these changes correspond to the ␧→␨ transition. Qualitatively, our data and those presented in Refs. 21 and 33 show similar trends, but detailed comparison shows different Raman spectra for the high-pressure phase ͑Fig. 3͒. We be- lieve that the disagreement arises from the use of different experimental procedures and the nature of the high-pressure phase ͑or phases͒. In contrast to the experiments reported in Refs. 21 and 33, we changed pressure at low temperature. It is useful to note that when infrared spectra were measured in a manner similar to ours,33 the results from the two studies ͑ ͒ FIG. 4. Raman spectra of the nitrogen vibrons through the transition to the agree very well inset to Fig. 3 . The evidence that the prop- ␨ phase at room temperature. The inset shows the variation of spectra with erties of the high-pressure phase depend on the thermody- temperature. The arrow designates a new peak, that appears at the transition. namic path suggests that this phase is not thermodynamically The spectra are shifted vertically for clarity. stable ͑i.e., is metastable at the indicated P – T conditions͒. This is supported by observations of a large hysteresis of the transition at low temperatures.33 An alternative ͑but related͒ ␯ rise to a new Raman peak ( 2b), indicated by an arrow in explanation is that the different properties of the high- Fig. 4 ͑see also Ref. 32͒. A similar observation was made pressure phase arise from relatively large pressure inhomo- when cooling down from approximately the same starting geneities in our experiment ͑since we changed pressure at point ͑inset in Fig. 4͒. At this pressure ͑and room tempera- low temperature͒. This is indicated by broadening of Raman ture͒ we also observed an increase in intensity of the infrared and infrared bands at higher pressures ͑Figs. 1 and 2͒. vibron.3 Raman spectra in the lattice mode range show split- The changes in the Raman and infrared spectra above 60 tings of the characteristic bands of the low-pressure phase GPa at room temperature are very similar to those observed and appearance of new low-frequency bands. The splitting of at low temperatures. Moreover, the reported transition the lattice modes increases with pressure, with the multiplets boundary33 extrapolated to high pressure and temperature evolving into distinct bands. Representative Raman spectra matches this room-temperature point. According to the ob- in the lattice-mode region are presented in Fig. 5. The com- served Raman and infrared spectra, the vibrational properties plexity of the spectra suggests a large number of molecules of the high-pressure phase are very similar at room and and low symmetry of the unit cell. The spectra at 95 GPa low temperature. Thus, we will consider it to be the same agree well with those presented in Ref. 32 in terms of the phase ͑␨͒. number and positions of the bands, with the exception of a In view of the absence of sufficient x-ray data for the ␨ Ϫ very weak additional vibron band near 2473 nm 1. With fur- phase, we can only speculate on its crystal structure. The ther pressure increase, the intensity of the lower-frequency number of vibron modes ͑in either our experiment or those vibron increases and all other vibron peaks gradually reported in Ref. 32͒ exceeds that predicted for R3c structure vanish.3,22 Similarly, in the lattice-mode region the intensity based on the space group theory proposed in Ref. 21. Ac- Ϫ of the 500 nm 1 band increases prior to the transition to the cording to Ref. 32, the increase of number of vibron modes is due to the increase in the number of different site symme- tries occupied by N2 molecules. Following this idea, up to 5 different site symmetry positions should be invoked to ex- plain the observed number of Raman vibron peaks above 60 GPa, which does not seem plausible. A critical examination of the spectra of Ref. 32 shows that this number can prob- ably be reduced to 3 according to the number of observed distinct peaks in the Raman exitations of the guest mol- ecules. Thus it seems natural to propose that the branching of vibron modes is related to sequential lifting of the degen- ␯ ␦ ͑␦ eracy of the 2 term of the cubic phase. In the first stage ␧ ͒ ␯ to transition , the 2 band splits into A1g and Eg compo- nents by the crystal field. In the second one ͑␧ to ␨ transi- tion͒, the symmetry is further reduced ͑to orthorhombic or FIG. 5. Raman spectra of nitrogen at 95 and 117 GPa. The inset shows a ͒ weak vibron peak ͑indicated by arrow͒ that develops under high pressure. monoclinic , with doubly degenerate level splitting into two The spectra are shifted in vertically for clarity. singlets. Additional splitting ͑vibrational or Davydov-type͒ Low Temp. Phys. 27 (99–10), September–October 2001 Goncharov et al. 869 of these major components could be caused by intermolecu- 8 R. M. Martin and R. J. Needs, Phys. Rev. B 34, 5082 ͑1986͒. lar interactions. This is related to a possible increase in the 9 C. Mailhiot, L. H. Yang, and A. K. McMahan, Phys. Rev. B 46, 14419 ͑ ͒ number of molecules in the unit cell as well as associated 1992 . 10 J. Belak, R. LeSar, and R. D. Etters, J. Chem. Phys. 92, 5430 ͑1990͒. symmetry lowering. High-quality diffraction data are re- 11 S. Nose´ and M. L. Klein, Phys. Rev. Lett. 50, 1207 ͑1983͒. quired in order to examine these hypotheses. 12 R. D. Etters, V. Chandrasekharan, E. Uzan, and K. Kobashi, Phys. Rev. B 33, 8615 ͑1986͒. 13 ͑ ͒ 5. CONCLUSIONS C. A. Swenson, J. Chem. Phys. 23, 1963 1955 . 14 R. L. Mills and A. F. Schuch, Phys. Rev. Lett. 23,1154͑1969͒. We find that the properties of the high-pressure, low- 15 F. Schuch and R. L. Mills, J. Chem. Phys. 52,6000͑1970͒. 16 ͑ ͒ temperature phase of nitrogen obtained by ‘‘cold’’ compres- J. R. Brookeman and T. A. Scott, J. Low Temp. Phys. 12,491 1973 . 17 W. E. Streib, T. H. Jordan, and W. N. Lipscomb, J. Chem. Phys. 37,2962 sion are different from those for the phase quenched from ͑1962͒. high temperature. This suggests that the ␨ phase is metastable 18 R. LeSar, S. A. Ekberg, L. H. Jones, R. L. Mills, L. A. Schwalbe, and D. Schiferl, Solid State Commun. 32, 131 ͑1979͒. and/or transitions to it are sensitive to nonhydrostatic effects. 19 ␧ ␨ S. Buchsbaum, R. L. Mills, and D. Schiferl, J. Phys. Chem. 88, 2522 Raman and infrared spectra of -N2 above 40 GPa and -N2 ͑1984͒. are not compatible with the R¯3c and R3c symmetries pro- 20 D. T. Cromer, R. L. Mills, D. Schiferl, and L. A. Schwalbe, Acta Crystal- posed in Ref. 21 because the number of vibron bands is logr., Sect. B: Struct. Crystallogr. Cryst. Chem. 37,8͑1981͒. 21 D. Schiferl, S. Buchsbaum, and R. L. Mills, J. Phys. Chem. 89, 2324 larger than predicted for the standard structures based on ͑1985͒. these space groups. This increase in the number of bands is 22 R. Reichlin, D. Schiferl, S. Martin, C. Vanderborgh, and R. L. Mills, Phys. probably related to additional lowering of the symmetry and Rev. Lett. 55, 1464 ͑1985͒. 23 multiplication of the size of the unit cell. R. L. Mills, B. Olinger, and D. T. Cromer, J. Chem. Phys. 84,2837͑1986͒. 24 S. Zinn, D. Schiferl, and M. F. Nicol, J. Chem. Phys. 87, 1267 ͑1987͒. The present vibrational spectroscopy data provide addi- 25 A. P. Jephcoat, R. J. Hemley, H.-K. Mao, and D. E. Cox, Bull. Am. Phys. tional constraints on the structure and properties of the high- Soc. 33, 522 ͑1988͒. pressure phases. They also suggest that known phases are not 26 H. Olijnyk, J. Chem. Phys. 93, 8968 ͑1990͒. 27 necessarily thermodynamically stable in the P – T region in H. Schneider, W. Haefner, A. Wokaun, and H. Olijnyk, J. Chem. Phys. 96, 8046 ͑1992͒. which they can be observed. As for other molecular crystals, 28 M. I. M. Scheerboom and J. A. Schouten, Phys. Rev. Lett. 71,2252 sluggish kinetics can complicate the determination of the ͑1993͒. true thermodynamic phase diagram ͑see, e.g., Ref. 39͒. Fur- 29 M. I. M. Scheerboom and J. A. Schouten, J. Chem. Phys. 105, 2553 ͑ ͒ ther theoretical and experimental effort is necessary to obtain 1996 . 30 M. Hanfland, M. Lorenzen, C. Wassilew-Reul, and F. Zontone, in Ab- a better understanding of the phase diagram of nitrogen at stracts of the International Conference on High Pressure Science and these high-pressure conditions. Technology, Kyoto, Japan, 1997, p. 130. 31 R. Bini, M. Jordan, L. Ulivi, and H. J. Jodl, J. Chem. Phys. 108, 6869 This work was supported by the NSF. The NSLS is sup- ͑1998͒. 32 ported by DOE. H. Olijnyk and A. P. Jephcoat, Phys. Rev. Lett. 83,332͑1999͒. 33 R. Bini, L. Ulivi, J. Kreutz, and H. J. Jodl, J. Chem. Phys. 112, 8522 ͑2000͒. * E-mail: [email protected] 34 V. G. Manzhelii and Y. A. Freiman, Physics of Cryocrystals, American Institute of Physics, College Park, MD ͑1997͒. 35 A. Mulder, J. P. J. Michels, and J. A. Schouten, J. Chem. Phys. 105, 3235 ͑1996͒. 1 R. J. Hemley and H.-K. Mao, in Encyclopedia of Applied Physics, edited 36 A. Mulder, J. P. J. Michels, and J. A. Schouten, Phys. Rev. B 57, 7571 by G. L. Trigg, New York, VCH Publishers, Vol. 18 ͑1997͒, p. 555. ͑1998͒. 2 V. Iota, C. S. Yoo, and H. Cynn, Science 283,1510͑1999͒. 37 A. F. Goncharov, V. V. Struzhkin, R. J. Hemley, H.-K. Mao, and Z. Liu, in 3 A. F. Goncharov, E. Gregoryanz, H.-K. Mao, Z. Liu, and R. J. Hemley, Science and Technology of High Pressure, M. H. Manghnani, W. J. Nellis, Phys. Rev. Lett. 85, 1262 ͑2000͒. and M. F. Nicol ͑Eds.͒, Universities Press, Hyderabad, India; Honolulu, 4 M. I. Eremets, R. J. Hemley, H. K. Mao, and E. Gregoryanz, Nature Hawaii ͑1999͒, Vol. 1, p. 90. ͑London͒ 411, 170 ͑2001͒. 38 Z. Liu, A. F. Goncharov, H. Mao, and R. Hemley, Eos Trans. AGU 81 5 E. Gregoryanz, A. F. Goncharov, R. J. Hemley, and H.-K. Mao, Phys. Rev. ͑48͒, Fall Meet. Suppl., Abstract T71B-22 ͑2000͒. B 64, 52103 ͑2001͒. 39 R. Jeanloz, J. Geophys. Res. 92, 10352 ͑1987͒. 6 E. Katoh, H. Yamawaki, H. Fujihisa, M. Sakashita, and K. Aoki, Phys. Rev. B 61,119͑2000͒. This article was published in English in the original Russian journal. Repro- 7 A. K. McMahan and R. LeSar, Phys. Rev. Lett. 54,1929͑1985͒. duced here with sylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Lattice modes of solid nitrogen to 104 GPa H. Olijnyk* and A. P. Jephcoat

Department of Earth Sciences, University of Oxford, Oxford OX1 3PR, United Kingdom ͑Submitted May 4, 2001͒ Fiz. Nizk. Temp. 27, 1175–1178 ͑September–October 2001͒ Lattice modes of solid nitrogen are studied by Raman spectroscopy at room temperature to 104 GPa using the diamond-anvil technique. Changes in the lattice mode spectral features correlate with those observed in the vibronic spectra suggesting symmetry changes of the crystal lattice. The changes in the spectral features mainly appear as branchings of existing modes, supporting the view of a close structural relationship among these high-pressure phases. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1414579͔

1. INTRODUCTION man spectra. This procedure also allows the determination of the frequencies of such modes, which manifest themselves as Solid nitrogen exhibits a complicated P – T phase dia- asymmetries or shoulders in the spectra. gram with a variety of different phases, which appears to be well established in the pressure range up to Ϸ20 GPa.1–22 Above Ϸ20 GPa solid nitrogen is less well understood, both 3. RESULTS from an experimental and theoretical point of view. Room Typical Raman spectra of the lattice-mode region at vari- temperature x-ray diffraction patterns between 16 and Ϸ60 ous pressures are shown in Fig. 1. One can note that the ¯ ͑⑀ ͒ 14,16,21 GPa are compatible with the R3c structure -phase . Raman spectra become increasingly more complex due to Raman studies at low temperatures and higher 10,11,18,22–25 the appearance of additional modes as pressure is increased. pressures suggest several structural modifications The mode frequencies and their pressure shifts are shown in Ϸ 26 above 25 GPa. Theoretical investigations have predicted Fig. 2. The previous room temperature results of Schneider a tetragonal lattice as the stable structure above 20 GPa, in 18 25 14,16,21 et al. and Goncharov et al. are also plotted. There is good disagreement with the experimental results, and above agreement with the data of Ref. 18 in the common pressure 12 GPa the calculated pressure–volume relations27 signifi- 16 range to 54 GPa, which indicates that any effects on the cantly depart from the accurate equations of state. Since the lattice modes due to the presence of the dilute isotopic spe- existence and properties of all these phases are determined cies can be neglected in the present study. In the pressure by the nature of the intermolecular interactions, these results demonstrate that improvements of the interaction potentials are necessary to correctly describe the properties of solid nitrogen above 20 GPa. In this context, the importance of the anisotropic part of the interaction potential for high-pressure structures of solid nitrogen has been pointed out recently.28,29 It was shown that the anisotropic term may be important in stabilizing the ␧-phase and also influences the orientational behavior of the molecules.28,29 A stringent test of the inter- molecular interaction potentials is provided by the lattice modes and their pressure dependences. In this paper we re- port Raman measurements of the external modes for pres- sures up to 104 GPa at room temperature.

2. EXPERIMENTAL 14 15 The sample, N2 enriched with 3% N2 and 1.4% 14N15N, was the same for which Raman studies of the internal modes were recently reported.23 The isotopic mix- ture was loaded into a high-pressure diamond-anvil cell at 0.2 GPa using a gas-loading technique.30 Raman spectra were excited by the 647 nm Krϩ laser line. Scattered light, collected through a spatial filtering aperture, was analyzed at an angle of 135° with respect to the incoming laser beam using a 0.6-m triple spectrograph and a liquid-nitrogen- cooled CCD multichannel detector. Pressure was determined with the hydrostatic ruby fluorescence scale.31,32 The peak positions were determined by fitting Voigt profiles to the Ra- FIG. 1. Lattice mode Raman spectra of solid nitrogen at various pressures.

1063-777X/2001/27(9–10)/3/$20.00870 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 H. Olijnyk and A. P. Jephcoat 871

TABLE I. Pressure coefficients and mode Gru¨neisen parameters of lattice modes of solid nitrogen for the pressure range from 16 to 104 GPa.

␯ Ϫ1 ␦ Ϫ1 ␦Ј Ϫ2 ␥ Mode 0 ,cm 0 , GPa 0 , GPa ␯ Ϫ L1a 249.9 0.05457 0.00781 — ␯ Ϫ L1b(1) 141.2 0.2201 0.124 1.51 ␯ Ϫ L1b(2) 130.5 0.3191 0.2780 1.51 ␯ Ϫ L8a 157.3 0.109 0.02956 1.59 ␯ ␯ Ϫ L8 , L8b 90.87 0.3242 0.2481 1.71 ␯ Ϫ L2a(1) 97.40 0.1777 0.06896 1.72 ␯ Ϫ L2a(2) 79.98 0.3564 0.3062 1.72 ␯ Ϫ L2b(1) 126.0 0.09057 0.01887 1.83 ␯ Ϫ L2b(2) 119.2 0.1479 0.0599 1.83 ␯ Ϫ L3 91.25 0.1409 0.04678 1.59 ␯ Ϫ L4 69.23 0.1606 0.05489 1.79 ␯ ␯ Ϫ L5 , L5a 37.87 0.3404 0.2039 2.23 ␯ Ϫ L5b 98.62 0.04155 0.00338 1.54 ␯ Ϫ L6a 46.20 0.08627 0.01102 — ␯ ␯ Ϫ L6 , L6b 59.10 0.05706 0.00626 1.62 ␯ Ϫ L7 55.52 0.02308 0.0003884 2.04

4. DISCUSSION The available x-ray diffraction data14,16,21 are compatible with the R¯3c lattice above 16.3 GPa and indicate a phase FIG. 2. Pressure shift of lattice mode frequencies of solid nitrogen. Solid transition at around 60 GPa by clear changes in the x-ray circles: present study; open triangles: Ref. 18. The thick gray lines represent diffraction pattern.14 The R¯3c lattice of the ␧-phase is a Ϸ the lattice mode data from Ref. 25, which were recorded to 150 GPa. slight distortion of the cubic ␦-phase and has two inequiva- lent sites, which correspond to the disklike and spherelike site of the ␦-phase, respectively.12 In the lattice-mode region range between 50 and 100 GPa, Goncharov et al.25 have re- eight Raman-active modes are allowed for R¯3c, two of ported six external Raman modes, whereas in the present which are librons and originate from the former spherelike study more than these six modes could be resolved in this site, whereas the remaining six modes are of translational pressure range, as can be noticed in Figs. 1 and 2. All modes character and are associated with the disklike site.10 show a positive frequency shift as pressure is increased. In The observed branchings or appearance of new lattice comparison to the earlier high-pressure Raman studies the modes around 30, 40, 65, and 80 GPa correlate well with the following additional observations have been made in the branchings observed for the internal modes of the host11,18,23 ␯ ␯ 23 present investigation. New modes, L6a and L1a , appeared as well as of the dilute isotopic species, as can be noted ␯ →␯ around 60 GPa, and further splittings, L2a L2a(1) , from the summary of Table II. ␯ ␯ →␯ ␯ L2a(2) and L1b L1b(1) , L1b(2) can be observed around The branchings in the internal modes occur only in the ␯ ␯ 80 GPa. With the exception of L7 , which becomes difficult 2 branch, which is related to the molecules on the disklike to observe above 60 GPa, all modes already present in the site. The splittings observed in the dilute isotopic species are ␧-phase above 16 GPa are still existent at the maximum pres- thought to result from inequivalent sites, since factor-group ␯ sure. Above 80 GPa the mode L4 , which merges into mode interactions, an alternative explanation for the splitting, are ␯ 23 L5a around 25 GPa, reappeared a few wave numbers below switched off at these small concentrations. ␯ ␯ ␯ mode L5a , indicating crossing of modes L5a and L4 Above 30 GPa, with nine lattice modes present, the Ra- around 55 GPa. The frequency–pressure data are well repre- man spectra of the external modes are no longer compatible sented by the expression with R¯3c, as are the vibronic spectra, which contrasts with Ϫ␦2 ␦Ј the lack of change observed in the x-ray diffraction patterns ␯͑ ͒ ␯ ϭ͓ Ϫ͑␦Ј ␦ ͒ ͔ 0/ 0 ͑ ͒ P / 0 1 0/ 0 P , 1 ␯ ϭ ␦ where 0 is the mode frequency at P 0 GPa, 0 ϭ ␯ TABLE II. Summary of branchings of lattice and vibronic modes in solid (d ln /dP)Pϭ0 is the logarithmic pressure coefficient, and ␦Ј ␦ ϭ N2. The number of inequivalent sites is determined from the splittings of the 0 is the pressure derivative of for P 0. The pressure isotopic guest vibrons.23 coefficients for the lattice modes are collected in Table 1. ␥ Lattice modes: Host Dilute guest Number of We have determined the mode Gru¨neisen parameters i ϭϪd ln ␯ /d ln V for the lattice modes. The V(P) values Branching Ref. 18, vibrons vibrons inequivalent i pressure present study ͑Ref. 23͒ ͑Ref. 23͒ sites were obtained from the P – V data of Refs. 14 and 16 using a Ϸ ␯ ␯ ␯ ␯ Birch-Murnaghan equation of state with bulk modulus B0 30 GPa L5b 2a 2a , 2b 3 Ϸ ␯ ␯ ␯ ␯ у ϭ2.55 GPa, its pressure derivative BЈϭ3.97, and V 40 GPa L2b , L8a 2c(2c) 2c2 4 0 0 Ϸ ␯ ␯ ␯ ␯ у ϭ 3 ␥ 65 GPa L6a , L1a 2c(2b) 2c1 5 50.99 Å /molecule. The average values of the i’s thus Ϸ ␯ ␯ 80 GPa L1b(1,2) , L2a(1,2) , splitting of obtained for the pressure range 16–65 GPa are also given in ␯ ␯ ␯ у L2b(1,2) 2c(1a) 2c2 6 Table I. 872 Low Temp. Phys. 27 (99–10), September–October 2001 H. Olijnyk and A. P. Jephcoat

The persistence of the low-pressure modes to the highest pressures, also observed for the internal modes, supports the argument of a close structural relationship among these high- pressure phases. The present data on lattice modes provide additional constraints, which have to be considered in future theoretical attempts at a quantitative interpretation of highly compressed solid nitrogen.

*E-mail: [email protected]

1 W. E. Streib, T. H. Jordan, and W. N. Lipscomb, J. Chem. Phys. 37, 2962 ͑1962͒. 2 R. L. Mills and A. F. Schuch, Phys. Rev. Lett. 23, 1154 ͑1969͒. 3 F. Schuch and R. L. Mills, J. Chem. Phys. 52, 6000 ͑1970͒. 4 J. A. Venables and C. A. English, Acta Crystallogr., Sect. B: Struct. Crys- ͑ ͒ FIG. 3. Thick gray solid lines are the calculated Raman-active lattice-mode tallogr. Cryst. Chem. 30, 929 1974 . 5 R. LeSar, S. A. Ekberg, L. H. Jones, R. L. Mills, L. A. Schwalbe, and frequencies for the R¯3c phase from Ref. 27. The solid triangles ͑Ref. 18͒ ͑ ͒ ͑ ͒ D. Schiferl, Solid State Commun. 32, 131 1979 . and solid circles present study are the experimentally observed frequen- 6 D. T. Cromer, R. L. Mills, D. Schiferl, and L. A. Schwalbe, Acta Crystal- cies. logr., Sect. B: Struct. Crystallogr. Cryst. Chem. 37,8͑1981͒. 7 D. Schiferl, D. T. Cromer, R. R. Ryan, A. C. Larson, R. LeSar, and R. L. Mills, Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 39,1151͑1983͒. in this pressure range. These changes in the spectral features 8 B. M. Powell, G. Dolling, and H. F. Nieman, J. Chem. Phys. 79,982 ͑ ͒ might be due to slight modifications in the lattice symmetry, 1983 . 9 S. Buchsbaum, R. L. Mills, and D. Schiferl, J. Phys. Chem. 88, 2522 possibly due to subtle changes in the orientation of mol- ͑1984͒. ecules at the disklike site, which may lead to two sites with 10 D. Schiferl, S. Buchsbaum, and R. L. Mills, J. Phys. Chem. 89, 2324 slightly different symmetry. Similar arguments may also be ͑1985͒. 11 valid for the 40 GPa branching. Another possibility might be R. Reichlin, D. Schiferl, S. Martin, C. Vanderborgh, and R. L. Mills, Phys. Rev. Lett. 55, 1464 ͑1985͒. that these branchings indicate the onset of the phase transi- 12 R. L. Mills, B. Olinger, and D. T. Cromer, J. Chem. Phys. 84,2837͑1986͒. tion, which has been observed by clear changes in the x-ray 13 S. Zinn, D. Schiferl, and M. F. Nicol, J. Chem. Phys. 87, 1267 ͑1987͒. diffraction pattern14 around 60 GPa and which would then 14 A. P. Jephcoat, R. J. Hemley, H.-K. Mao, and D. E. Cox, Bull. Am. Phys. ͑ ͒ occur gradually over a large pressure range. In this connec- Soc. 33, 522 1988 . 15 D. Schiferl, R. LeSar, and D. S. Moore, in Simple Molecular Systems at tion it is interesting to note that the boundary between the Very High Density, edited by A. Polian, P. Loubeyre, and N. Boccara, ␧-phase and the higher-pressure ␨-phase,24 extrapolated to Plenum Press, New York ͑1989͒, p. 303 and references therein. room temperature, implies an ␧→␨ transition at 50–60 GPa. 16 H. Olijnyk, J. Chem. Phys. 93, 8968 ͑1990͒. 17 B. J. Baer and M. Nicol, High Press. Res. 4,511͑1990͒. The 80 GPa branching has also been interpreted as an indi- 18 H. Schneider, W. Haefner, A. Wokaun, and H. Olijnyk, J. Chem. Phys. 96, cation of a further phase transition, which is supported by the 8046 ͑1992͒. ␯ 15 19 observed splitting of the 2c2 mode of the N2 isotopic spe- M. I. M. Scheerboom and J. A. Schouten, Phys. Rev. Lett. 71,2252 cies. One characteristic feature common to both the external ͑1993͒. 20 M. I. M. Scheerboom and J. A. Schouten, J. Chem. Phys. 105, 2553 and internal modes is the persistence of the lower-pressure ͑1996͒. modes at the highest pressures, which implies a close struc- 21 M. Hanfland, M. Lorenzen, C. Wassilew-Reul, and F. Zontone, in Ab- tural relationship among the various high-pressure phases. stracts of the International Conference on High Pressure Science and The observed lattice-mode frequencies are compared Technology, Kyoto, Japan, 1997, p. 130. 22 R. Bini, M. Jordan, L. Ulivi, and H. J. Jodl, J. Chem. Phys. 108, 6869 27 with calculated frequencies of the R¯3c lattice to 40 GPa in ͑1998͒. Fig. 3. The calculated frequencies of modes ␯ , ␯ , and 23 H. Olijnyk and A. P. Jephcoat, Phys. Rev. Lett. 83,332͑1999͒. L1 L3 24 ␯ are in very good agreement with experiment, whereas for R. Bini, L. Ulivi, J. Kreutz, and H. J. Jodl, J. Chem. Phys. 112, 8522 L4 ͑2000͒. the other modes the good agreement found at lower 25 A. F. Goncharov, E. Gregoryanz, H.-K. Mao, Z. Liu, and R. J. Hemley, 27 pressures gets lost at higher pressures. Possible reasons for Phys. Rev. Lett. 85, 1262 ͑2000͒. this increasing discrepancy might involve the inadequacy, at 26 J. Belak, R. LeSar, and R. D. Etters, J. Chem. Phys. 92, 5430 ͑1990͒. 27 higher pressures, of the interaction potential used in these R. D. Etters, V. Chandrasekharan, E. Uzan, and K. Kobashi, Phys. Rev. B 33, 8615 ͑1986͒. calculations, a different crystal structure above 30 GPa, and 28 A. Mulder, J. P. J. Michels, and J. A. Schouten, J. Chem. Phys. 105, 3235 finite temperature effects. ͑1996͒. 29 A. Mulder, J. P. J. Michels, and J. A. Schouten, Phys. Rev. B 57, 7571 ͑1998͒. 5. CONCLUSION 30 A. P. Jephcoat, H.-K. Mao, and P. M. Bell, in Hydrothermal Experimental Techniques, edited by G. C. Ulmer and H. L. Barnes, Wiley Interscience, The Raman spectra of the lattice modes of solid N2 are New York ͑1987͒, p. 469. compatible with the R¯3c structure between 16.3 GPa and 31 R. A. Forman, G. J. Piermarini, J. D. Barnett, and S. Block, Science 176, ͑ ͒ Ϸ30 GPa. At higher pressures ͑30, 40, 65, and 80 284 1972 . 32 H.-K. Mao, J. Xu, and P. M. Bell, J. Geophys. Res. 91, 4673 ͑1986͒. GPa͒changes in the spectral features of the lattice modes correlate with changes in the spectral features of the internal This article was published in English in the original Russian journal. Repro- modes, probably related to changes in the lattice symmetry. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Real-space condensation in a dilute Bose gas at low temperature I. O. Kulik*

Department of Physics, Bilkent University, Bilkent, Ankara 06533, Turkey ͑Submitted May 16, 2001͒ Fiz. Nizk. Temp. 27, 1179–1182 ͑September–October 2001͒ We show with a direct numerical analysis that a dilute Bose gas in an external potential—which is chosen for simplicity as a radial parabolic well—undergoes at a certain temperature Tc a phase transition to a state supporting a macroscopic fraction of particles at the origin of the phase space ͑rϭ0, pϭ0͒. Quantization of particle motion in a well wipes out the sharp transition but supports a distribution of a radial particle density ␳(r) peaked at rϭ0 ͑a real-space condensate͒ as well as a phase-space Wigner distribution density W(r,p) peaked at rϭ0 ϭ and p 0 below a crossover temperature Tc* of order of Tc . A fixed-particle-number canonical ensemble, which is a combination of the fixed-N condensate part and the fixed-␮ excitation part, is suggested to resolve the difficulty of large fluctuation of the particle number (␦NϳN)in the Bose-Einstein condensation problem treated within the orthodox grand canonical ensemble formalism. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1414580͔

The phenomenon of Bose-Einstein ͑BE͒ condensation then the solution for the equilibrium state will be ͑ ͒ see textbooks, e.g., Refs. 1–3 manifests itself in the forma- Ϫ Ϫ ␶ 1 f 0ϩ␶ 1 f 0 tion of macroscopic fraction of zero-momentum particles ϭ 1 1 2 2 ͑ ͒ f ␶Ϫ1ϩ␶Ϫ1 . 3 uniformly distributed in a coordinate space. Such transition 1 2 was recently observed in laser-trapped, evaporation-cooled ␶Ϫ1 atomic vapors4–6 in magnetic traps ͑see recent reviews7–9͒. The relaxation rate 2 is proportional to the laser intensity ␶Ϫ1ӷ␶Ϫ1 ͑ ͒ We will show by a direct numerical analysis, partly similar to P. At large intensity, assuming 2 1 , Eq. 3 gives f Ӎ 0 and sometimes overlapping with the previous theoretical f 2. works on the subject,10–13 that a Bose gas in an external In a semiclassical approximation, the particle energy is confining potential condenses at low temperature to a posi- p2 1 tion of minimum potential energy; the particles of that ‘‘con- ␧ϭ ϩ m⍀2r2, ͑4͒ 2m 2 densate’’ also have zero kinetic energy. Quantization of par- ticle states in a well makes the real-space condensation a where the thermodynamic potential ⍀ϭϪT ln Z, Z is the continuous transition rather than a phase transition but still grand partition function ͑assuming zero spin of particles͒ supports a macroscopic fraction of particles near the origin of the coordinate space below a crossover temperature T* dpdr c ϭ ͵ ͑ Ϫ ͑␮Ϫ␧͒/T͒ ͑ ͒ Z 3 ln 1 e , 5 which is of the order of Bose-condensation temperature Tc . ͑2␲ប͒ Experimental realization of BE condensation implies ប ␮ confinement of a dilute gas within some region of space in a where is Planck’s constant. The chemical potential is ͑ ͒ ‘‘trap’’ cooled by its interaction with an ‘‘optical molasses’’ determined from 5 to satisfy an equation 14 created by laser irradiation and finally cooled to dpdr 1 microwave-range temperature by evaporative cooling.11 ϭ ͵ ͑ ͒ N ͑ ␲ប͒3 ͑␧Ϫ␮͒/TϪ , 6 Bose gas in a trap may be considered to be interacting with 2 e 1 two thermal reservoirs, the first one representing the thermal where N is the number of particles. After integration over the ͑ ͒ environment walls, blackbody radiation at temperature T1 directions of r and p we obtain and the second one the optical molasses at temperature T2 ӶT . The equilibrium distribution of particles f (p, r, t) can ͑4␲͒2 2T 3/2 1 Nϭ ͑2mT͒3/2ͩ ͪ be obtained by solving the Boltzmann kinetic equation ͑2␲ប͒3 m⍀2

df ϱ ϱ ϭˆ ͕ ͖ϩˆ ͕ ͖ ͑ ͒ 1 I1 f I2 f , 1 ϫ ͵ 2 ͵ 2 ͑ ͒ dt x dx y dy 2ϩ 2Ϫ␨ , 7 0 0 ex y Ϫ1 where Iˆ is the interaction term ͑Stoss integral͒ correspond- 1 where ␨Ͻ0 is the chemical potential in appropriate dimen- ˆ ing to coupling with a media 1, and I2, respectively, with sionless units. media 2. If we choose for simplicity the relaxation time ap- At low temperature, no nonzero value of ␨ can satisfy proximation for Iˆ , ͑ ͒ ϭ 1,2 Eq. 7 . It therefore vanishes at a temperature T Tc0 deter- Ϫ mined from the condition ␨ϭ0, thus giving f f i Iˆ ϭϪ , ͑2͒ i ␶ ϭប⍀ ␨ ͒͒1/3ϭ ប⍀ 1/3 ͑ ͒ i Tc0 ͑N/ ͑3 0.94 N , 8

1063-777X/2001/27(9–10)/3/$20.00873 © 2001 American Institute of Physics 874 Low Temp. Phys. 27 (99–10), September–October 2001 I. O. Kulik

FIG. 1. Chemical potential versus temperature for various values of N:102 FIG. 2. Radial density distribution ␳(r)ϭ4␲r2n(r) for Nϭ1000 and vari- ͑ ͒ 3 ͑ ͒ 4 ͑ ͒ 5 ͑ ͒ 6 ͑ ͒ ϭ ͑ ͒ ͑ ͒ ͑ ͒ ͑ ͒ 1 ,10 2 ,10 3 ,10 4 ,10 5 . ous temperatures: T/Tc0 0.2 1 , 0.8 2 ,1.4 3 , 2.0 4 .

␨ ␨ Figure 2 shows the radial density distribution ␳(r) where (z) is the Riemann zeta function. Below Tc0 , re- ϭ4␲r2n(r) at various temperatures. Below T* , ␳(r) displays mains equal to zero with the total number of particles N0 c having both rϭ0 and pϭ0 values, determined from a second maximum at small r, which grows in amplitude as the temperature decreases: the real-space condensate. The T3 ϭͩ Ϫ ͪ ͑ ͒ formation of such a condensate is even more explicit in the N0 1 3 N. 9 Tc0 evolution of the z-projected density distribution, Fig. 3, as the temperature is decreased from above to below T . ϭ ϭ c0 Of course, the r 0, p 0 state is not allowed quantum- At zero temperature, all excited particles above the con- ͑ ͒ ͑ ͒ mechanically, and the derivation leading to Eqs. 6 , 7 densate vanish. The joint momentum–coordinate distribution needs modification. The energy of a particle in a parabolic function ͑the Wigner distribution function16͒ takes a value well, Eq. ͑4͒,is N0 Ϫ 2 2 Ϫ 2 2 ␧ϭប⍀͑ ϩ ϩ ϩ ͒ ϭ ͑ ͒ϭ p r0 r /r0 ͑ ͒ n1 n2 n3 3/2 , ni 0,1... . W p,r ␲ e e , 13 r0 ͑ ͒ Then the normalization condition, Eq. 6 , reduces to ϭ ប ⍀ 1/2 where r0 ( /m ) is the zero-point oscillation amplitude ϱ S in a parabolic well. ϭ n ͑ ͒ N ͚ nx 10 The question remains, how to reconcile the above results nϭ0 ␩e Ϫ1 with the free-space Bose-Einstein condensation. The BE con- with densation temperature equals1 n 1 ប2 ϭ ␦ ϭ ϩ ͒ ϩ ͒ ϭ 2/3 ͑ ͒ Sn ͚ n ϩn ϩn ,n ͑n 1 ͑n 2 T0 3.31 n . 14 ϭ 1 2 3 m n1 ,n2 ,n3 0 2 ␩ϭ ␮ Ϫ␮ ϭប⍀ ␮ The average density of particles in a well above the con- and exp(( 0 )/T), x /T; 0 is the value of the ϭ ␮ ϭ ប⍀ densation temperature is chemical potential at T 0( 0 3/2 ). The solution of Eq. ͑10͒ shows the dependence ␮(T) T 1/2 ͑ ͒ ¯nϳN/T3, where ¯rϭͩ ͪ ϳr N1/6͑T/T ͒1/2, ͑15͒ Fig. 1 with a crossover between almost linear dependence m⍀ 0 0 above the crossover temperature Tc* and a practically zero value below that temperature. The value of Tc* is very near to ӷ Tc0 at large number of particles, N 1. The particle density distribution is expressed through the sum of Hermite polynomials.15 Employing the identity for these polynomials r H2 ͑x ͒ nk k ͚ ͟ n ϩ ϩ ϩ ϭ ϭ k n1 n2 ... nr n k 1 2 nk! n 1 r 1/2 ϭ 2 ͩͩ 2 ͪ ͪ ͑ ͒ ͚ rnϪm m Hm ͚ xk , 11 mϭ0 2 m! kϭ1 ϭ ϭ where rm 1 for m even and rm 0 for m odd, we receive by putting rϭ3

2 ϱ ϱ eϪr H2 ͑r͒ 1 ͑ ͒ϭ m ͑ ͒ FIG. 3. Side view of particle distribution: 1—Tϭ0.2T , 2—Tϭ0.8T , n r 3/2 ͚ m ͚ ͑ ϩ ͒ . 12 c0 c0 ␲ ϭ ϭ ␩ m 2k xϪ ϭ ϭ m 0 2 m! k 0 e 1 3—T 1.4Tc0 , 4—T 2.0Tc0 . Low Temp. Phys. 27 (99–10), September–October 2001 I. O. Kulik 875 where ¯r is a confinement radius ͑mean radius of the gaseous ӍeN. This agrees with the conclusion, reached in a different ͒ cloud . It is related to the minimal quantum radius r0 as ¯r way in Ref. 12, that the thermodynamic properties of a Bose ϳ 1/6 1/2 ϳ r0N (T/Tc) . By putting T Tc* as defined above, we condensate in a trap with fixed total number of particles are obtain T of the order of the BE condensation temperature very similar to those in the orthodox grand canonical en- ͑14͒. Therefore, the phenomenon we discussed is just the BE semble with a fixed average number of particles. The above condensation mechanism,1 except that in a trap the conden- results are consistent with a known statement that the Bose- sation occurs in both the momentum and coordinate spaces Einstein condensation temperature T0 is the same in the ca- or, if we choose to explore the behavior of a dilute low- nonical and in the grand canonical ensembles.2 temperature Bose gas in real space, it will condense there, In conclusion, I hope I have met the goal of elucidating making up a high-density globular fraction coexisting with in a direct way the properties of the low-temperature state of the spatially dispersed ‘‘excitations’’ in a region of size com- an ideal Bose gas of finite-size, finite-particle-number sys- parable to the thermal confinement radius ¯r. tems. I express my deep gratitude to Prof. B. Tanatar for In the grand canonical ensemble which we so far have stimulating discussions and help. been considering, the number of particles is not fixed. The mean square fluctuation of particle number in a state ␣ is ␦ 2 ϭ ϩ *Also at: B. Verkin Institute for Low Temperature Physics and Engineering ͗ n␣͘ n␣(n␣ 1). In a condensate, by putting ͗n␣ϭ0͘ ϭ ␮Ӎ␧ Ϫ ͗␦ 2͘1/2Ӎ of the National Academy of Sciences of Ukraine, 47 Lenin. Ave., N0 we get 0 T/N0 and n0 N0 . This means a Kharkov 61103, Ukraine. ␦ ϳ Ӷ huge fluctuation of particle number N N at T T0 ,an E-mail: [email protected] unrealistic property of the model.17 In a canonical ensemble, which better fits to experiments with dilute gases in traps, the average value of the conden- 1 sate population is given by L. D. Landau and E. M. Lifshitz, Statistical Physics, Vol. 1, Pergamon, New York ͑1987͒. Ϫ␤͚ ͑␧ Ϫ␧ ͒ 2 ͑ ͒ ͚N ͚ ␣Ͼ0 ␣ 0 n␣␦ K. Huang, Statistical Mechanics, Wiley, New York 1987 . ϭ n0 ͕n ͖Јe ͚ n , NϪn 3 n0 0 ␣ ␣Ͼ0 ␣ 0 A. H. Carter, Classical and Statistical Thermodynamics, Prentice Hall, ͗n ͘ϭ Ϫ␤͚ ͑␧ Ϫ␧ ͒ , ͑16͒ 0 ͚N ͚ ␣Ͼ0 ␣ 0 n␣␦ New Jersey ͑2001͒. ϭ ͕n ͖Јe ͚ n , NϪn n0 0 ␣ ␣Ͼ0 ␣ 0 4 M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E. Wieman, and E. A. ͕ ͖Ј Cornell, Science 269, 198 ͑1995͒. where n␣ stands for a collection of all state numbers ex- 5 ␤ϭ K. B. Davis, M. O. Mewes, M. A. Joffe, M. R. Andrews, and W. Keterle, cept n0 , and 1/T. The average over such states does not Phys. Rev. Lett. 74, 5202 ͑1995͒. fluctuate strongly and therefore can be replaced by its grand 6 W. Petrich, M. H. Anderson, J. R. Ensher, and E. A. Cornell, Phys. Rev. canonical value corresponding to an appropriate choice of Lett. 74, 3352 ͑1995͒. 7 F. Dalfovo, S. Giorgini, L. P. Pitaevskii, and S. Stringati, Rev. Mod. Phys. chemical potential ␮ϭ␮ Ϫ . We thus get N n0 71,463͑1999͒. 8 ͚N W. Ketterle, D. S. Dufree, and D. M. Stamper-Kurn, Making, Probing and n ϭ0 n0ZNϪn ͗ ͘Х 0 0 ͑ ͒ Understanding Bose-Einstein Condensates, Preprint xxx.lanl.gov/abs/ n0 ͚N , 17 cond-mat/9904034 ͑1999͒. ϭ Z Ϫ 9 n0 0 N n0 E. A. Cornell, J. R. Ensler, and C. E. Wieman, Experiments in Dilute Ϫ␤⍀ Atomic Bose-Einstein Condensation, Preprint xxx.lanl.gov/abs/cond-mat/ ϭ n ⍀ where Zn e , and n is the thermodynamic potential of 9903109 ͑1999͒. 1 the grand canonical ensemble. 10 N. L. Balazs and T. Bergeman, Phys. Rev. A 58, 2359 ͑1998͒. ϭ Ϫ␤N 11 ͑ ͒ The quantity Zn e is not exponentially small for a W. Ketterle and N. J. van Druten, Phys. Rev. A 54, 656 1996 . 12 P. Navez, D. Bitouk, M. Gajda, Z. Idziaszek, and K. Rzazewski, Phys. number of particles n smaller than the Bose-condensate frac- ͑ ͒ Ͻ ͑ ͒ Rev. Lett. 79, 1789 1997 . tion, n N0 . Therefore, we can change expression 17 to 13 F. Brosens, J. T. Devreese, and L. F. Lemmens, Phys. Rev. E 55,6795 Ϫ␤⍀ ͑1997͒. ͚N NϪn n ϭN n0e 0 14 ͗ ͘Х 0 0 ͑ ͒ Y. Castin, J. Dalibard, and C. Cohen-Tannoudji, Laser Cooling and Trap- n0 N Ϫ␤⍀ Ϫ . 18 ping of Neutral Atoms,inAtoms in Electromagnetic Fields, edited by C. ͚ ϭ e N n0 n0 N0 Cohen-Tannoudji, World Scientific, Singapore ͑1994͒. 15 ⍀ ϭ I. S. Gradsteyn and I. M. Ryzhik, Tables of Integrals, Sums and Products, The quantity n is strongly peaked at n N0 , thus giving edited by A. Jeffrey, Acad. Press, New York ͑1994͒. ͗ ͘Ӎ ͗␦ 2͘1/2ϳͱ 16 n0 N0 and, similarly, n0 N0 rather than E. P. Wigner, Phys. Rev. 40, 749 ͑1932͒. ͗␦ 2͘1/2ϳ 17 M. Fierz, Helv. Phys. Acta 29,47͑1995͒. n0 N0 as in the orthodox grand canonical ensemble. Ӷ ͑ ӷ ͒ Indeed, at N N0 corresponding to T T0 we obtain for the This article was published in English in the original Russian journal. Repro- ⍀ ⍀ ӍϪ thermodynamic potential N a value N NT, and ZN duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Linear and nonlinear waves on the charged surface of liquid hydrogen M. Yu. Brazhnikov, G. V. Kolmakov, A. A. Levchenko,* and L. P. Mezhov-Deglin

Institute of Solid State Physics, 142432 Chernogolovka, Moscow District, Russia ͑Submitted June 22, 2001͒ Fiz. Nizk. Temp. 27, 1183–1191 ͑September–October 2001͒ The results of research on the properties of linear and nonlinear waves on the charged surface of liquid hydrogen in a cylindrical cell are reported. It is found that the spectrum of oscillations of linear waves softens with increasing applied electric field. Weak turbulence in a system of capillary waves on the charged surface of liquid hydrogen is investigated. The formation of a Kolmogorov cascade is observed in the inertial interval from 100 Hz to 10 kHz. It is found that the correlation function of the deviation of the surface from its flat equilibrium state can be described by a power-law function of the frequency, with an exponent mϭϪ3.7Ϯ0.3 when the surface is excited at a single resonance frequency, and mϭϪ3.0Ϯ0.3 in the case of two- frequency excitation. The results of these studies are in qualitative agreement with the theoretical predictions. © 2001 American Institute of Physics. ͓DOI: ͓DOI: 10.1063/1.1414581͔

INTRODUCTION on the surface of liquid hydrogen by an external force which is many times smaller than in the case of an ordinary liquid Research on the dynamics of waves and turbulence on such as water. This circumstance is decisive for the use of a the surface and in the bulk of a liquid is of great interest both technique in which the waves on the surface are excited by for basic nonlinear physics and from the standpoint of prac- electrical forces. Previous experiments have shown that the tical applications. One approach to the description of the surface of liquid hydrogen can be charged by injecting turbulent state has been proposed in the theory of weak tur- charges into the bulk of the liquid, that the charges can be bulence ͑see Ref. 1 and the references cited therein͒, which held near the surface for a long time, and that surface waves in the case of capillary and gravity waves on the free surface can be excited by an alternating electric field.4,5 An advan- of a liquid predicts a power-law dependence of the distribu- tage of this technique is that the external force can act di- tion of the energy of the waves over their frequency—an rectly on the surface of the liquid, almost completely exclud- isotropic Kolmogorov spectrum. The predictions of the ing the bulk, and also the high degree of isotropicity of the theory of weak turbulence are in good agreement with the exciting force, which makes it possible to study turbulence results of a numerical simulation. Meanwhile, despite an ap- under well-controlled experimental conditions. preciable number of studies on the nonlinear dynamics of surface waves, at the present time there are only a few pub- lished reports of experimental observations of isotropic spec- 2. THEORETICAL MODEL tra on the surface of a liquid. For example, in Refs. 2 and 3 2.1. Spectrum of small oscillations of the charged surface the transition to a regime of weak capillary turbulence on of a liquid increasing amplitude of the surface waves was observed at frequencies up to 1 kHz. The spectrum of linear oscillations of the charged sur- In this paper we report a study of linear and nonlinear face of a liquid found in a gravity field between horizontal capillary waves on the charged surface of liquid hydrogen. plates of a flat capacitor is described by the known 6 The spectrum of small oscillations of the surface of the liq- expression uid and its evolution with increasing surface charge density ␴k2 2kP ␻2ϭ ͑ ͒ͫ ϩ Ϫ ͑ ͒ͬ ͑ ͒ are investigated up to the critical value at which the flat k k tanh kh g ␳ ␳ coth kd , 1 surface loses stability. The formation of a Kolmogorov cas- ␻ cade in the system of capillary waves on the surface of liquid where k is the frequency of a wave with wave vector k, h hydrogen was observed at frequencies up to 10 kHz when is the thickness of the liquid layer, ␴ is the coefficient of the surface was excited at one or two frequencies. surface tension, d is the distance from the surface of the We describe a technique of excitation of waves on the liquid to the upper plate of the capacitor, Pϭ(U/d)2/8␲ is charged surface of liquid hydrogen and of measuring the the pressure exerted by the electrical forces on the flat, un- frequency dependence of the correlation function of the de- disturbed surface, U is the potential difference across the viations of the surface from equilibrium. capacitor plates, g is the acceleration of free fall, and ␳ is the The advantages of liquid hydrogen in experiments on density of the liquid. It is assumed in this formula that the nonlinear dynamics include the relatively low kinematic vis- electric field is completely screened by surface charges. In cosity and the large coefficient of nonlinearity of capillary the limiting case when the distance from the surface of the waves, making it possible to broaden significantly the fre- liquid to the control plate of the capacitor is small, so that quency band in which a turbulent cascade is formed. In ad- kdӶ1, and the depth is large, khӷ1, the spectrum of the dition, because of the low density, oscillations can be excited charged surface of a liquid can be written in the form

1063-777X/2001/27(9–10)/7/$20.00876 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 Brazhnikov et al. 877

␴k2 flux in the direction of higher frequencies and, consequently, ␻2ϭ ͩ ϩ ͪ ͑ ͒ k k G ␳ , 2 is realized at frequencies above the pump frequency ͑direct cascade͒. where the quantity GϭgϪ2P/␳d plays the role of an effec- The stationary distribution of the energy of surface tive acceleration of free fall. At voltages near the critical waves in the inertial interval can be described by a pair cor- ϭ ͉␩ ͉2 value at which the surface loses stability for the case of a thin relation function in the Fourier representation, I␻ ͗ ␻ ͘, ϭ ␲␳ 3 1/2 ͑ ͒ for the deviations ␩(r,t) of the surface from the flat state: film, Uc1 (4 gd ) Ref. 6 , the effective acceleration ␻ Ϫ17/6 G goes to zero, and k becomes a monotonic power-law I␻ϭconst ␻ . ͑5͒ function of the wave vector with exponent 3/2, starting with • small k: The energy distribution is related to the pair correlation function by the relation ␻ Х͑␴ ␳͒1/2 3/2 ͑ ͒ k / k . 3 E␻ϭ␳gI␻ . This circumstance is important for choosing the lowest The theoretical prediction is confirmed by the results of pump frequency in studying turbulence. In our experiments numerical calculations of the nonlinear evolution of capillary in the low-frequency region the conditions kdϽ1, khϽ1 waves directly from first principles on the basis of the equa- 9 hold, and the minimum wave vector above which relation ͑3͒ tions of hydrodynamics. is valid must be estimated experimentally. In the case of spectrally narrow pumping a numerical 10 For a deep liquid khӷ1 and a large distance to the con- calculation shows that a system of equidistant peaks at trol electrode kdӷ1, as the voltage U is increased, relation multiples of the pump frequency arises on the I␻ curve. The ͑1͒ develops a local minimum at a value of the wave vector frequency dependence of the height of the peaks is described of the order of the inverse capillary length 1/␭. At a critical by a power-law function with an exponent of Ϫ21/6. This ϭ ␲␴␭Ϫ1/2 1/2 ␻ value differs by 2/3 from the exponent in Eq. ͑5͒, which was value Uc (2 ) d the frequency k goes to zero. This means that the flat charged surface loses stability at the obtained for the case of pumping in a wide frequency inter- critical voltage.7 val. We note that the power-law dependence on frequency for 2.2. Weak turbulence in a system of capillary waves the correlation function of the deviations of the surface from Turbulence is the name given to a highly excited state of equilibrium has been observed at frequencies up to 1 kHz in a system with many degrees of freedom, the system being far experiments on water,2,3 from measurements of the power from its thermodynamic equilibrium.1 The turbulent regime spectrum of radiation transmitted through a layer of liquid is characterized by a strong nonlinear interaction of the de- whose surface was excited at a low frequency by means of a grees of freedom and by significant energy dissipation. Tur- vibrating platform. It was shown that for observation of a bulence can be observed in systems where the excitation power-law distribution of energy over frequency it is neces- ͑pump͒ frequencies and the frequency at which strong energy sary that the amplitude of the wave at the pump frequency be dissipation occurs are widely separated on the frequency higher than a certain critical value ͑around 1 mm͒.Inthe scale. The nonlinear interaction leads to an efficient redistri- experiments of Ref. 2 the exponent in the correlation func- bution of energy among the degrees of freedom ͑modes͒. tion was equal to Ϫ3/2, and in Ref. 3 the exponent was close The main question in the study of turbulence is to find to the theoretical value Ϫ17/6. the energy distribution law over the frequency scale, Our preliminary results for the charged surface of liquid E␻—the stationary spectrum of the turbulence. hydrogen were published in Ref. 11, where we showed that An ensemble of interacting waves can be described in in the frequency interval 100–5000 Hz the function I␻ can the framework of a kinetic equation analogous to the Boltz- be described by a power-law function with an exponent of mann equation of gas dynamics. The dispersion relation for Ϫ3Ϯ0.5. capillary waves, ␻ϭ(␴/␳)1/2k3/2 is of the decay type, and, consequently, the main contribution to the interaction of 3. METHODS OF MEASUREMENT AND PROCESSING OF waves is from three-wave processes—the decay of a wave RESULTS into two with conservation of the total wave vector and total The scheme of the measurements of the spectrum of sur- frequency, and also the inverse of this process, the conflu- face oscillations is shown in Fig. 1. The experiments were ence of two waves into one. In the system of capillary waves done in an optical cell placed in a helium cryostat. Inside the there exists a frequency interval ͑inertial interval͒ bounded ␻ cell was a horizontal flat capacitor. A radioactive plate was below by the pump frequency p and at high frequencies by the viscous damping, in which a power-law dependence of the energy distribution over the frequency of the waves is observed. The theory of uniform capillary turbulence8 pre- dicts an energy distribution law ͑Kolmogorov spectrum͒ of the form

1/2 Ϫ17/6 E␻ϭcp ␻ , ͑4͒ where p is the energy flux in k space, and c is a constant. The distribution ͑4͒ is characterized by a constant energy FIG. 1. Scheme of the measurements of the spectrum of surface oscillations. 878 Low Temp. Phys. 27 (99–10), September–October 2001 Brazhnikov et al. placed on the lower capacitor plate, which was made in the spond to the situation when the excitation frequency coin- form of a disk 25 mm in diameter. Hydrogen was condensed cides with one of the resonance frequencies of the surface into a cup formed by the lower plate and a guard ring 25 mm oscillations. This makes it possible to determine the discrete in diameter and 2.7 mm in height. The upper plate of the spectrum of eigenfrequencies of the oscillations of the sur- ␻ ϭ capacitor, situated above structure surface of the liquid, was face of the liquid in the experimental cell, n f (n), where made in the form of a cone with a flat base 25 mm in diam- n is the number of the resonance. eter. The distance from the liquid to the upper plate was 2.5 Under conditions of a cylindrical geometry a standing mm in the experiments on measurement of the dispersion wave on the surface of the liquid is described by the equation ␻(k) and 4 mm for the study of turbulence. The temperature ␨͑ ͒ϭ ͑ ͒ ͑␻ ͒ ͑ ͒ of the liquid in the experiments was around 16 K. r,t AnJ0 knR cos nt , 6 The free surface of the liquid was charged by means of a where J0 is the Bessel function of order zero, R is the inner radioactive plate emitting ␤ electrons. The electrons ionized radius of the guard ring, and An is the amplitude of the wave. a thin layer of the liquid near the plate. A dc voltage was The relation between the number of the resonance and the applied across the capacitor plates. The sign of the charges ϭ wave vector of the resonant standing wave kn s(n) is found forming a quasi-two-dimensional layer beneath the surface from the equation of the liquid was determined by the polarity of the voltage. ͑ ͒ϭ In the experiments reported here the oscillations of a posi- J1 knR 0. tively charged surface were investigated. A metal guard ring This is equivalent to the condition that there are 2n nodes of mounted around the radioactive plate prevented the charges the standing wave along the diameter 2R of the cell. from escaping beneath the surface to the walls of the con- Thus, knowledge of the functions ␻ ϭ f (n) and k tainer, and it can therefore be assumed that the density of n n ϭs(n) enables one to uniquely determine the dispersion re- charges in the layer is proportional to the applied voltage and lation ␻ ϭ␻(k ) of the surface waves. that the electric field in the bulk of the liquid is close to zero. n n In the experiments on turbulence the beam reflected The oscillations of the surface of liquid hydrogen were from the oscillating surface of the liquid was focused on the excited by an ac voltage applied to the guard ring in addition photodetector by the lens. Thus in that experiment the total to the dc voltage. The amplitude of the ac voltage was much power of the beam was measured. The angle between the less than the dc voltage. The oscillations of the surface of the beam and the undisturbed flat surface of the liquid ͑the graz- liquid were registered with the aid of a laser beam reflected ing angle͒ was ␣ϭ0.2 rad. The laser beam lay in a vertical from the surface. plane passing through a diameter of the cell. In the experi- We used two schemes for registering the oscillations of ments we used a narrow and a wide laser beam. The light the liquid surface. In the experiments on measurement of the spot on the surface of the liquid was an ellipse with axes dispersion relation ␻(k) the frequency of the ac voltage ap- a ϭ0.5 mm and b ϭ0.1 mm in the case of the narrow beam, plied to the guard ring was varied smoothly from 1.5 to 150 1 1 or a ϭ2.5 mm and b ϭ0.5 mm for the wide beam. The out- Hz. At certain frequencies a resonant standing wave was es- 2 2 put signal of the photodetector, which was directly propor- tablished on the liquid hydrogen surface. The amplitude of tional to the total power P(t) of the beam, was stored in a the oscillations of the laser beam reflected from the surface computer with the aid of a 12-bit analog-to digital converter increased sharply and became larger than the dimensions of for several seconds at a frequency of 25 KHz. the photodetector ͑the collecting lens shown in Fig. 1 was We analyzed the frequency spectrum P␻ of the total removed in this scheme͒, and so the average value of the power of the reflected laser beam, obtained by Fourier trans- light intensity registered by the photodetector decreased sig- forming the P(t) data with respect to time. nificantly. To improve the accuracy of the measurements the The total power of the laser beam depends on the graz- signal from the photodetector was averaged in time. Thus the ing angle and on the relationship between the wavelength ␭ minima on the curve describing the frequency dependence of on the surface of the liquid and the linear dimensions of the the voltage registered on the photodetector ͑Fig. 2͒ corre- light spot. In our experiments the amplitude of the oscilla- tions of the surface at the pump frequency was small, so that the angle of deviation of the oscillating surface from the flat state, ␸р10Ϫ2 rad, was much less than the grazing angle ␣. For small-amplitude waves with wavelengths much greater than the size of the light spot, ␭ӷa, the power of the reflected beam is a linear function of the angle ␸.12 The angle ␸ can be estimated as the ratio of the amplitude of the wave to its wavelength. Then in the frequency representation the correlation function of the deviations of the surface from equilibrium ͑the ‘‘low-frequency limit’’͒ is written as

2 Ϫ2/3 2 2 Ϫ4/3 I␻ϭ͉͗␩␻͉ ͘ϳ͑␻ ␸␻͒ ϳP␻␻ . ͑7͒ In the opposite case, ␭Ͻa, when several wavelengths FIG. 2. Experimental trace of the average intensity of light as a function of lay within the dimensions of the spot, the registered power of the frequency of the ac voltage applied to the guard ring. The numbers of the standing-wave resonances on the hydrogen surface are indicated near the the reflected beam is determined by the value of the angle of minima. The dc voltage was Uϭ954 V. inclination of the surface averaged over the area of the light Low Temp. Phys. 27 (99–10), September–October 2001 Brazhnikov et al. 879

spot. A calculation shows that the change in power of the reflected beam on the whole is proportional to the product of the amplitude of the variation of the angle ␸ and the wave- ␭ ϳ␭␸ ͑ length , i.e., P␻ w . This leads to the relation the ‘‘high-frequency limit’’͒ 2 I␻ϳP␻. ͑8͒ The frequency of a wave on the surface of liquid hydro- gen with a wavelength ␭Ϸa, near which one expects a crossover from the ‘‘low-frequency’’ regime of registering the oscillations to the ‘‘high-frequency’’ regime, is equal to ␻ ␲Ϸ ␻ ␲Ϸ FIG. 4. Total power P of the narrow laser beam reflected from the charged l/2 500 Hz for the narrow beam and l/2 50 Hz for the wide beam. hydrogen surface as a function of time t. The pumping frequency was 27.5 Hz.

4. EXPERIMENTAL RESULTS tance from the surface to the upper plate dϭ2.5, which were 4.1. Linear waves of the order of the capillary length. Relation ͑1͒ gives a good Figure 2 shows a typical experimental trace of the aver- description of the experimental data in the entire interval of aged light intensity registered by the photodetector as a func- voltages. With increasing voltage the spectrum of the surface tion of the frequency ␻/2␲ of the ac voltage applied to the waves softens—the frequency of the oscillations with a guard ring. The deep minima on the curve correspond to the given wave vector kn decreases. The dotted line corresponds formation of standing waves on the surface. The number of to the power-law function ␻ϳk3/2, which describes the dis- the resonance is given near the minima. The Q factor of the persion of capillary waves on the surface of a deep liquid. experimental cell in the investigated frequency interval was We see that at voltages above 0.8 Uc1 the dispersion relation ␻ Ϸ 30–50, permitting reliable measurement of the resonance n(kn) of the surface waves under the conditions h d frequency ␻ . From the experimental traces of the intensity Ϸ Ϫ1 n kc is well approximated by this function. like that shown in Fig. 2, we found the relation between the The inset shows the spectrum of oscillations of the number of the resonance and the frequency of the ac voltage, charged surface of liquid hydrogen calculated according to ␻ ϭ n f (n). From this function and the wave vectors calcu- formula ͑1͒ for conditions when the depth of the liquid and ϭ lated from the function kn s(n), we recovered the spectrum the distance from the charged surface to the control electrode ␻ of oscillations of the surface, n(kn). The results of a pro- are much greater than the wavelengths (khӷ1, kdӷ1, i.e., cessing of the experimental data obtained at three different tanh(kh)ϭ1, coth(kd)ϭ1) for three different voltages U less voltages between the capacitor plates, Uϭ283, 954, and than the critical value ͑unlike the case of the experimental 1080 V, are shown by the data points in Fig. 3 on a log–log plots, here a linear scale is used͒. We see that as the voltage scale. The temperature of the measurements was T U is increased, a local minimum develops on the ␻(k) curve ϭ16.0 K. The measured critical voltage, at which recon- in the region of wave vectors of the order of the inverse ϭ ␳ ␴ 1/2Ϸ Ϫ1 struction of the flat charged surface of the liquid occurs, was capillary length kc ( g/ ) 5cm . A similar spectrum ϭ Uc1 1140 V, i.e., measurements were made both at volt- with a minimum was observed previously in experiments ages much less than the critical value Uc1 and also at volt- with helium in Ref. 7. ages close to Uc1 . Notice the qualitative difference of the spectra in the two The solid lines in Fig. 3 were calculated according to limiting cases: the spectrum of oscillations of the surface of ͑ ͒ Ϸ ␭ relation 1 for voltages of 283 and 1080 V and the specified a thin layer does not exhibit the local minimum at kc 1/ values of the liquid layer thickness hϭ2.7 mm and the dis- that is observed for the deep liquid.

4.2. Nonlinear waves We investigated the frequency distribution of the ampli- tude of the deviations ␩(r,t) of the surface of the liquid hydrogen from equilibrium for excitation of the surface by a harmonic force at frequencies from 25 to 300 Hz ͑at reso- nances with numbers 3 and higher͒ for different levels of pumping. Figure 4 shows an experimental trace of the time depen- dence P(t) of the total power of a narrow laser beam re- flected from the surface. The surface was excited at a fre- ␻ ␲ϭ quency p/2 27.5 Hz. The maximum angle of deviation of the surface from equilibrium at the pump frequency was 0.03 rad, and the maximum amplitude of the wave was of the order of 0.1 mm. The frequency spectrum P␻ of the total FIG. 3. Dispersion curves for waves on the charged surface of a liquid hydrogen layer 2.7 mm thick. The inset shows the evolution of the disper- power was obtained by Fourier transforming the measured sion curve ␻(k) on the charged surface of a deep liquid. dependence P(t). Figure 5 shows the square of the Fourier 880 Low Temp. Phys. 27 (99–10), September–October 2001 Brazhnikov et al.

tained in the experiment with the wide beam. The measure- ments were made under the same conditions as in the experi- ment with the narrow beam: pump frequency 27.5 Hz, dc voltage Uϭ1170 V, amplitude of the wave at the pump fre- quency around 0.1 mm. We recall that an estimate of the frequency of the crossover from the regime of registration of long-wavelength oscillations to the regime of registration of short-wavelength oscillations gave a value of 50 Hz. Thus, according to Eq. ͑8͒, over almost all of the frequency range the correlation function is directly proportional to the square of the amplitude of the Fourier harmonic of the measured total power of the reflected laser beam. We see that the varia- FIG. 5. Distribution of the square of the Fourier harmonics of the total tion of the height of the peaks with frequency on the graph of power of a narrow laser beam reflected from the liquid hydrogen surface. 2 ␻Ϫ3.7 The pumping frequency was 27.5 Hz. P␻ can be described by a power law , starting at 300 Hz and on up to 7 kHz. The difference of the exponent from Ϫ17/6 agrees qualitatively with the conclusion10 that in the 2 2 amplitude P␻ as a function of frequency ␻/2␲. The P␻ case of a spectrally narrow pump the exponent increases in ␻ curve has a main peak at the pump frequency p and peaks absolute value in comparison with the value of the exponent at multiples of that frequency. These peaks correspond to in the case of pumping in a wide frequency interval. capillary waves which are generated on the surface of the In addition, in our experiments the spectrum of surface liquid due to nonlinearity. At frequencies up to 800 Hz the ␻ oscillations (kn) is discrete, and the width of the reso- frequency dependence of the height of the peaks can be de- nances is much less than the distance between them. This 2 ϳ␻Ϫ1.3 scribed by a power-law function of the type P␻ .At may be an additional circumstance leading to the difference ␻ ␲Ͼ high frequencies /2 800 Hz the frequency dependence in the values of the observed exponents of the power laws of the height of the peaks is described by a stronger power- from the value predicted by the theory,8,10 since the calcula- Ϫ law function with an exponent close to 3.7. At frequencies tions were done for systems with a quasicontinuous spectrum above 4 kHz the peak disappears into the instrument noise. of oscillations. The change of the exponent of the power-law function One notices the strong difference of the frequency de- describing the frequency dependence of the peak height ob- 2 pendence of P␻ at low frequencies for the cases of the nar- served near 800 Hz can be attributed to a transition from the row ͑Fig. 5͒ and wide ͑Fig. 6͒ laser beams. In Fig. 6 the regime of registration of long-wavelength oscillations with height of the peaks of P2 in the frequency interval from 27.5 ␻Ͻ␻ ␻ l to the regime of registration of short-wavelength os- to 300 Hz is not described by the power-law function of ␻Ͼ␻ cillations ( l). The observed value of the crossover fre- frequency of the form ␻m. ␻ ␲ϭ quency l/2 800 Hz is close to the value of the crossover The differences in the spectra shown in Figs. 5 and 6 are ϳ frequency estimated in Sec. 3 ( 500 Hz). due solely to a change in the diameter of the laser beam and, ͑ ͒ ͑ ͒ Using formulas 7 and 8 , we find that the correlation consequently, relate only to the condition for the onset of the function of the deviations of the surface at frequencies below 2 crossover regime. The nonmonotonicity of P␻ can be attrib- 800 Hz is proportional to ␻Ϫ2.6, while at high frequencies Ϫ3.7 uted to a feature of the optical method used to detect the I␻ϳ␻ . surface oscillations.13 One of the reasons for this, as our Thus the experimental data demonstrate a power-law fre- observations show, may be that in the spectrum P␻ of the quency dependence of the correlation function of the devia- registered signal the amplitudes of the Fourier harmonics at tions of the surface ͑scaling͒ in the frequency interval 100– low frequencies depend on the position of the laser spot on 4000 Hz, but with different values of the exponent at low and the surface of the liquid. At the same time, the high- high frequencies. frequency part of the spectrum does not depend on the posi- Figure 6 shows the frequency dependence of P2 ob- ␻ tion of the laser beam, since the size of the spot is much greater than the wavelengths. It should be noted that increasing the amplitude and fre- quency of the pump leads to a broadening of the frequency interval in which power-law frequency dependence of the 2 correlation function is observed. Figure 7 shows plots of P␻ obtained for excitation of the surface at 135 Hz for two dif- ferent pumping levels, corresponding to standing-wave am- plitudes of 0.004 and 0.03 mm at the excitation frequency. 2 The sharp drop in the height of the peaks on P␻ at frequen- cies from 2 to 3.5 kHz in Fig. 7a may be due to viscous damping of the oscillations at the edge of the inertial inter- val. As expected,1 when the amplitude of the exciting force is increased, the high-frequency edge of the inertial interval FIG. 6. Distribution of the square of the Fourier harmonics of the total shifts to higher frequencies. We can conclude from Fig. 7b power of a wide laser beam. The pumping frequency was 27.5 Hz. that the frequency of this edge exceeds 10 kHz. Low Temp. Phys. 27 (99–10), September–October 2001 Brazhnikov et al. 881

FIG. 8. Distribution of the square of the Fourier harmonics of the total power of a wide laser beam for pumping at frequencies of 27.5 and 81 Hz.

␻Ϫ3Ϯ0.3. This differs from the dependence ␻Ϫ3.7Ϯ0.3 ob- served in the experiments with the excitation of the surface at a single frequency. The decrease of the absolute value of the exponent from 3.7 to 3 may be due to the change of the conditions of excitation of the waves on the surface. We note that the value mϭϪ3 is close to the theoretical estimate m ϭϪ17/6 obtained for the case of pumping in a wide fre- FIG. 7. Evolution of the distribution of the square of the Fourier harmonics quency interval.8 of the total power of a wide laser beam with increasing pumping level. The Figure 9 shows another spectrum of excitation of oscil- amplitude of the standing wave at the pumping frequency, 135 Hz, is 0.004 ␻ ␲ϭ ͑ mm ͑a͒ and 0.03 mm ͑b͒. lations at two frequencies, 1/2 27.5 Hz the third reso- ͒ ␻ ␲ϭ ␻ nance and 2/2 45 Hz. Since the frequency 2 does not coincide with a resonance frequency of the cell, for observa- Thus one can assert that for excitation of the surface of tion of the combination frequencies it was necessary to in- ␻ liquid hydrogen in a cylindrical cell by a periodic external crease the amplitude of the ac voltage at the frequency 2 as force, when a standing wave with an amplitude of the order much as possible and to decrease the amplitude at the fre- ␻ of 0.1 mm arises on the surface, there exists a wide fre- quency 1 , so that the amplitude of the waves would be quency interval in which the correlation function of the de- comparable to each other. In the figure we see peaks corre- viations of the surface from equilibrium are proportional to a sponding to oscillations of the surface at the combination ␻ Ϫ␻ ␻ ϩ␻ power-law function of the frequency, with an exponent close frequencies 2 1 and 2 1 . The excitation frequencies Ϫ Ϯ ϳ␻Ϫ3.7Ϯ0.3 ␻ Ϫ␻ to 3.7 0.3, i.e., I␻ . were chosen such that the combination frequency 2 1 would coincide with the second resonance of the cell, 17.5 ␻ ϩ␻ 4.3. Combination frequencies Hz, and the combination frequency 2 1 would be found between the fifth and sixth resonances. As a result, the am- In order to elucidate the influence of the conditions of plitude of the peak at the frequency ␻ Ϫ␻ is almost 10 excitation of the surface and to mitigate the effect of the 2 1 times greater than the amplitude of the peak at the frequency choice of the position of the laser spot on the surface of the ␻ ϩ␻ . Consequently, the discrete character of the spec- liquid on the form of the low-frequency part of the registered 2 1 2 spectrum P␻ , we did a series of experiments in which the surface was excited simultaneously at two resonance fre- 2 quency of the cell. Figure 8 shows a plot of P␻ in the case of ␻ ␲ϭ ␻ ␲ pumping at the frequencies 1/2 27.5 Hz and 2/2 ϭ81 Hz ͑the third and seventh resonances͒. The pumping was done by two independent generators, so that the phase relation between the standing waves was arbitrary. The two ␻ ␻ main peaks at the frequencies 1 and 2 are clearly visible in the figure, and one can also see combination peaks corre- ␻ Ϫ␻ ␻ ϩ␻ sponding to frequencies 2 1 and 2 1 . For each peak at a frequency ␻ one can find a combination frequency ␻ ϭ ␻ Ϯ ␻ p 2 q 1 , where p and q are integers. With this change in the conditions of excitation, the dip in the frequency in- terval 27.5–200 Hz that is clearly seen in Fig. 6 has practi- cally vanished. FIG. 9. Frequency dependence of the square of the Fourier harmonics of the At frequencies above 100 Hz the frequency dependence total power of the wide laser beam for pumping at frequencies of 27.5 and 2 of the height of the peaks of the spectrum P␻ is close to 45 Hz. 882 Low Temp. Phys. 27 (99–10), September–October 2001 Brazhnikov et al. trum of eigenfrequencies of the oscillations of the liquid sur- The authors thank V. E. Zakharov and E. A. Kuznetsov face turns out to have a substantial influence on the depen- for helpful discussions, E. Henry, P. Alstrom, and M. 2 dence of P␻ at low frequencies. Levinsen for providing the text of Ref. 3 prior to its publi- cation, and V. N. Khlopinskii for assistance in preparing the 5. CONCLUSIONS experiments. This study was supported in part by the Minis- try of Industry, Science and Technology of the Russian Fed- The experimentally measured spectrum of linear oscilla- eration ͑project ‘‘Kristall-6’’͒ and by the grant INTAS- tions of the charged surface of liquid hydrogen in an external NETWORK 97-1643. electric field is in good agreement with the predictions of the theoretical calculations6 of the dispersion relation ͑1͒ of sur- face waves on an equipotential charged surface of a liquid *E-mail: [email protected] layer located between the plates of a flat capacitor. With increasing value of the external stretching field the spectrum of oscillations of the charged surface of liquid hydrogen soft- 1 V. Zakharov, V. L’vov, and G. Fal’kovich, Kolmogorov Spectra of Turbu- ens and approaches the dependence ␻ϳk3/2 characteristic for lence, Vol. 1, Wave Turbulence, Springer-Verlag, Berlin ͑1992͒. 2 capillary waves on the surface of a deep liquid. W. Wright, R. Hiller, and S. Putterman, J. Acoust. Soc. Am. 92,2360 ͑1992͒. In the experiments investigating nonlinear waves excited 3 E. Henry, P. Alstrom, and M. T. Levinsen, Europhys. Lett. 52,27͑2000͒. by a periodic external force, a powerlike dependence on fre- 4 A. A. Levchenko and L. P. Mezhov-Deglin, Fiz. Nizk. Temp. 22,210 quency ͑scaling͒ was observed for the correlation function of ͑1996͓͒Low Temp. Phys. 22, 162 ͑1996͔͒. 5 A. A. Levchenko, G. V. Kolmakov, L. P. Mezhov-Deglin, M. G. the deviations of the surface of the liquid hydrogen from ͑ ͒ Ϸ Mikhailov, and A. B. Trusov, J. Low Temp. Phys. 119,343 2000 . equilibrium at frequencies up to 10 kHz. For excitation of 6 D. M. Chernikova, Fiz. Nizk. Temp. 2, 1374 ͑1976͓͒Sov. J. Low Temp. the surface by a harmonic force at a fixed frequency, corre- Phys. 2, 669 ͑1976͔͒. sponding to one of the resonance frequencies of the cell, the 7 P. Leiderer, Phys. Rev. B 20, 4511 ͑1979͒. Ϫ3.7Ϯ0.3 8 V. E. Zakharov and N. N. Filonenko, Zh. Prikl. Mekh. Tekh. Fiz. 5,62 correlation function I␻ϳ␻ . ͑1967͒. For excitation of the surface simultaneously at two fre- 9 A. N. Pushkarev and V. E. Zakharov, Phys. Rev. Lett. 76, 3320 ͑1996͒. quencies, the correlation function depends on frequency as 10 G. E. Fal’kovich and A. B. Shafarenko, Zh. E´ ksp. Teor. Fiz. 94, 172 Ϫ3Ϯ0.3 ͑ ͓͒ ͑ ͔͒ I␻ϳ␻ . 1988 Sov. Phys. JETP 67, 1393 1988 . 11 M. Yu. Brazhnikov, A. A. Levchenko, G. V. Kolmakov, and L. P. Mezhov- Combination frequencies of the surface oscillations were Deglin, JETP Lett. 73, 398 ͑2001͒. observed and studied. It was shown that the discrete charac- 12 L. D. Landau and E. M. Lifshitz, Electrodynamics of Continuous Media, ter of the eigenfrequencies of the oscillations of the surface 2nd ed., rev. and enl., by E. M. Lifshitz and L. P. Pitaevskii, Pergamon in a cell of finite size has a substantial influence on the fre- Press, Oxford ͑1984͒, Nauka, Moscow ͑1982͒, p. 405. 13 M. Yu. Brazhnikov and A. A. Levchenko, Prib. Tekh. E´ ksper. ͑in press͒. quency distribution of the amplitudes of the oscillations at low frequencies. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Excitonic luminescence in the drift of excess electrons through liquid and solid rare gases E. B. Gordon*

Branch of the Institute of Energy Problems of Chemical Physics, Russian Academy of Sciences, 142432 Chernogolovka, Moscow District, Russia A. F. Shestakov**

Institute of Problems of Chemical Physics, Russian Academy of Sciences, 142432 Chernogolovka, Moscow District, Russia ͑Submitted June 25, 2001͒ Fiz. Nizk. Temp. 27, 1192–1201 ͑September–October 2001͒ It is shown that the excitation of electronic states is the main channel of energy loss of excess electrons drifting in moderate (у103 V/cm) electric fields through condensed heavy rare gases ͑Rg͒. These losses, together with scattering on resonances of metastable negative ions (Rg*)Ϫ, determine the dependence of the average energy of the electrons and their drift velocity vd on the electric field E both in condensed Rg and in dense gases. In particular, explanations are given for the constancy of vd at large E and for the transformation of the electroluminescence spectrum upon changes in the density of heavy particles and their temperature. Thus it is predicted that localized excitons can be efficiently excited in the bulk of crystalline and liquid Xe, Kr, and Ar, with a yield of around 102 excitons ͑and UV photons͒ per electron. © 2001 American Institute of Physics. ͓DOI: ͓10.1063/1.1414582͔

The elucidation of the properties of collective electronic Those hopes, as we know, were completely justified, excitations—excitons—in insulator and semiconductor crys- and, although the most widely used excimer lasers utilize tals is one of the universally recognized outstanding achieve- rare-gas halides, the first first report of light amplification on ments of the Kharkov scientific school co-founded by L. V. excimer molecules was in a paper devoted to the excitation Shubnikov.1 Particular attention was devoted to the study of of liquid xenon by an electron beam.3 rare-gas crystals, which served Frenkel as a first model for Since the characteristics of an exciton are largely deter- creating a theory of excitons.2 Three types of collective ex- mined by the properties of the electronically excited particle citations have been observed in rare gases ͑Rg͒: coherent generating it, in respect to its general features ͑position of the excitons, and one-center ͑quasi-atomic͒ and two-center spectral bands, transition probability͒ a quasi-atomic exciton ͑quasi-molecular͒ localized states. The last of these are bears a surprising resemblance to an excited rare gas atom mainly responsible for the luminescence of crystals of heavy Rg, and the molecular exciton, which we are mainly inter- ͑ ͒ ͑ ͒ rare gases Xe, Kr, and Ar excited by vacuum ultraviolet ested in, resembles the excimer molecule Rg2* Ref. 1 . The ͑VUV͒ radiation or by an electron beam. A distinguishing specifics of a solid are manifested in the dynamics of trans- feature of heavy rare-gas crystals is the extraordinarily high formation of an ‘‘atomic’’ excitation into a ‘‘molecular’’ ex- efficiency of conversion of the energy of photons or fast citation and in the features of the structure of the spectral electrons into excitonic luminescence occurring in the VUV bands. Naturally, investigators had a keen desire to create an region: ␭ϭ172 nm for Xe, ␭ϭ148 nm for Kr, and ␭ excitonic laser using condensed inert gases, primarily xenon, ϭ127 nm for Ar. but efforts to do so did not meet with much success,4 and for At that same time work was ongoing at the P. N. what we think is not a fundamental reason: the impossibility Lebedev Physics Institute ͑FIAN͒ to study the features of the of arranging a high-power pump. Indeed, pumping by the electronic excitation and the resulting luminescence in dense radiation from another UV laser5 was not of practical inter- rare gases. Those studies led to the discovery of a new object est, the use of a VUV lamp was inefficient, and fast electron in molecular physics—excimer molecules or exciplexes, beams, which are used to pump the dense-gas Xe2 excimer chemical compounds which are quite stable but exist only in laser,6 do not penetrate deeply enough into the volume of a an electronically excited state. The absence of a bound crystal or liquid. ground state automatically ensures a population inversion of At the same time, it happens by a chance coincidence the levels of the corresponding phototransition and a wide that in heavy rare gases, the same substances in which di- luminescence spectrum; this, together with another feature of atomic excimer molecules are characteristically formed in excimer molecules—a huge increase in the probability of an the gas phase ͑in He and Ne this is impossible for a number optical transition in comparison with the analogous transition of fundamental reasons͒,7 the mobility of excess electrons in in a united atom, immediately aroused interest in excimer the solid and liquid is quite high, almost the same as for molecules as a possible active medium for high-power UV conduction electrons in metals ͑for He and Ne it is low be- lasers. cause of the formation of so-called ‘‘bubbles’’͒.8 In principle,

1063-777X/2001/27(9–10)/7/$20.00883 © 2001 American Institute of Physics 884 Low Temp. Phys. 27 (99–10), September–October 2001 E. B. Gordon and A. F. Shestakov

the reasons for such a high mobility are clear: first, rare gases tures of the method of forming electronic excitations inside condense as close-packed atoms; second, the high values of rare-gas crystals on the basis of the drift of excess electrons the mass ratio of the Rg atom to the electron; third, the through them. For analysis we use a simple two-parameter absence of electron affinity among Rg atoms. model that is surely valid for a dense gas and which is often The drift of excess electrons in an electric field through used for describing the drift of excess electrons in condensed liquid and solid rare gases has been studied by many rare gases ͑the aforementioned nonspecificity of the behavior authors,9 and a number of hypotheses have been put forth to of the drift velocity with respect to the phase state serves as explain the observed effects, viz.: 1͒ the existence of a pla- a justification for this͒. teau in the dependence of the electron drift velocity vd on the We consider the drift of an electron as a process of its 15 applied field E at large E, and the essential lack of depen- random walk in the force field eE. We introduce the fol- ͒ ͑ ␭ dence of vd on the phase state of the rare gas; 2 a sharp by lowing parameters: m—the mean free path with respect to more than 3 order of magnitude in Xe͒ increase of the elec- total loss of momentum direction, so that after a ‘‘collision’’ ␭ tron mobility in low fields upon the transition of the gas to the average velocity of the electron is equal to zero; w—the the condensed state and the nonmonotonic character of the mean free path with respect to loss of energy; now ␧ ϵ␭ ␭ temperature dependence of the mobility in the condensed m / w is the fraction of the energy lost after one ‘‘colli- ␧Ӷ state, in combination with the absence of any features at the sion’’ accompanied by loss of momentum ( 1). solid–liquid phase transition. Another important question Neglecting the time of collision in comparison with the that has been discussed is whether ‘‘hot’’ electrons can exist mean free time, we can assume that the electron acquires in the drift of electrons through condensed rare gases; some energy from the electric field during the time of its free mo- authors have assumed that the velocity of the electrons inside tion between collisions. Then its energy after the ith collision a solid cannot exceed the speed of sound,10 while others have with loss of momentum is equal to assumed that electrons are heated to energies of several 2 ϩ ͒2 2 2 2 11 mviϩ1 m͑vi ati mvi ma ti electron-volts. ϭ͑1Ϫ␧͒ ϭ͑1Ϫ␧͒ͫ ϩ ͬ, Finally, in 1994, one of the present authors ͑E.B.G.͒ 2 2 2 2 12 proposed a model in which it was assumed that not only where m is the mass of the electron, aϭeE/m is its accel- can the excess electrons acquire high energy from the elec- eration in the field E, and ti is the time between the ith and tric field, all the way up to the first excitation potential of the (iϩ1)th collisions. Thus if we neglect the change in the total Rg atom, but this excitation is even the main channel of velocity over the time between two successive collisions and energy loss for an excess electron and determines its average the difference of the trajectory from rectilinear, i.e., ti energy and drift velocity. The electronically excited Rg atom ϭ␭ m /vi , the recursion relation for the kinetic energy of an formed on electron impact ͑a one-center exciton͒ rapidly ͑es- electron has the form pecially in the case of Xe͒ reacts with the neighboring atoms 2 ͑ W ϩ w and transforms into a two-atom excimer molecule two- i 1 ϭ͑ Ϫ␧͒ͫ ϩ 0 ͬ ͑ ͒ ͒ 1 Wi , 1 center exciton which upon its decay emits a VUV photon. 2 4Wi This means, in particular, that simple means ͑an electric field in an electrolytic cell͒ can be used to form excitonic excita- where w0 is the energy acquired by the electron from the tions in the bulk of a crystal or liquid, and one primary field over its mean free path: electron can generate a large number of excited centers. Fur- ␻ ϭ ␭ ͑ ͒ 0 eE m , 2 thermore, in a cell with a photocathode, on account of the appearance of secondary electrons upon their photoemission or, in differential form, for ␧Ӷ1, under the action of the excimeric VUV radiation, it is pos- d͑W2͒ sible for interesting critical phenomena to arise, in particular, ϭ␻2Ϫ ␧ 2 ͑ ͒ 2 0 4 W . 3 electron and, hence, photon avalanches. di The practical aspects of the problem are also important: In low electric fields the electrons are in thermal equilibrium if the model proposed in Ref. 12 is valid, it would mean that with the substance. Isothermal drift is characterized by a con- it is possible in principle to create an efficient direct con- ϭ stant mobility, since vd at0/2, and t0 , the mean free time of verter of electrical energy into VUV radiation, including the electron in its steady drift, is constant and equal to coherent.13 ␭ m /vT , where vT is the thermal speed of the electron. Thus Generally speaking, UV emission upon the application ϳ vd E and is given by the equation of an electric field in liquid rare gases has been observed ␭ experimentally, primarily in Xe; it was ascribed to a so- eE m w0 v ϭ ϭ . ͑4͒ called ‘‘glow discharge’’ and was not given a reliable expla- d 3/2 2mvT 2 ͱmkT nation. Only the recent direct experiments of Wyder’s group14 have confirmed the conclusions of our papers;12,13 in Expression ͑4͒, up to a numerical coefficient of 1.5, agrees particular, they recorded the intensity of VUV emission in with the formula obtained by Bardeen and Shockley10 for the the emission band of Xe2* in the drift of electrons emitted by electron mobility in nonpolar crystals on the basis of the a cold cathode in liquid xenon in moderate electric fields, Boltzmann equation. and they found that the influence of the emission depended For moderately high fields, when the electron energy linearly on the applied voltage. WӷkT, and under the condition of stationarity of Eq. ͑3͒ is In the present paper we analyze the advantages and fea- equal to Low Temp. Phys. 27 (99–10), September–October 2001 E. B. Gordon and A. F. Shestakov 885

w0 Wϭ , ͑5͒ 2ͱ␧ the gas approximation is even more realistic: even for the motion of an electron in the conduction band in a xenon crystal, the mass of a conduction electron is close to m,asa calculation shows.1 Condition ͑5͒ can be rewritten in the form ␭ 1 eE m at0 vϭ ͩ ͪͩ ͪ ϭ . ͑6͒ ͱ␧ m v 23/2ͱ␧ The drift velocity of an electron under these conditions is, as ϭ before, equal to vd at0/2. Hence v ͱ␧ FIG. 1. Electron drift velocity in Xe versus the electric field according to the d ϭ ϳͱ ͑ ͒ data of Ref. 17: 1—gas, nϭ6ϫ1019, Tϭ172 K; 2—liquid, Tϭ165 K; ; vd E. 7 v 2 3—solid, Tϭ155 K. The dashed lines have slopes of 1 and 1/2. For com- parison the electroluminescent yield of Xe at Pϭ0.35 bar and Tϭ300 K The physical meaning of expression ͑7͒ is obvious—in the from Ref. 20 is also shown ͑4͒. random walk of a particle its displacement is always deter- mined by the square root of the number of steps ͑in this case, to the loss of energy͒. For free atoms ͑a gas, the hard-spheres tron immediately loses all of its energy. Therefore, under approximation͒␧¯ ϭ2m/M, and expression ͑7͒ becomes condition ͑9͒ the kinetic energy of the drifting electron, hav- ing increased to I0 , then falls sharply back practically to vd ϭͱm/2M, ͑8͒ zero, and these cycles repeat during the whole course of the v drift, so that the average energy W and average velocity v of Хͱ the electron remain constant. The question arises, how will which is practically the same as the result vd /v m/M which is known from calculations and experiments on the the drift velocity vd of an excess electron behave as a func- drift of electrons in a gas ͑M is the mass of the Rg atom͒.16 tion of the applied electric field E? ͑ ͒ Thus in the drift of an electron in an electric field ͑with The approximate solution of equation 1 for the energy accuracy up to the dependence of the electron scattering of an excess electron Wi after the ith step of a random walk cross section on its energy on account of the Ramsauer ef- has the form fect, which is rather smooth and reduces to a parametric 2 w0 dependence ␧(E)͒ the linear dependence of the drift velocity W2ϭ ͑1ϪeϪ2i␧͒, ͑10͒ i 4␧ on the applied field, which is characteristic for thermal elec- trons, gives way at higher fields ͑when the electron tempera- and its average drift velocity is given by ture ‘‘detaches’’ from the temperature of the heavy particles͒ 2 i ati to a square-root dependence on E. i ͚ m i Ϫ ͚ m s iϭ0 ͚ m v 2 It is hard to expect that the hard-spheres approximation i iϭ0 i 2 w0 iϭ0 i v mϭ ϭ ϭ . ͑11͒ and, hence, formula ͑8͒ will be valid for a liquid and solid, d ͚im ͚im 2m ͚im Ϫ1 iϭ0ti iϭ0ti iϭ0vi particularly in view of the fact that at the densities charac- teristic for the condensed state the effective electron scatter- The step number im at which the energy of the electron ing cross section of Rg atoms is an order of magnitude reaches the energy of electronic excitation is determined ϭ smaller than in a dense gas.17,18 However, the results of from the equation Wi I0 and is equal to experiments17 and calculations19 show that the energy loss 1 ␧ i ϭϪ ln͑1Ϫ␣2͒. ͑12͒ per electron scattering event, , remains close to 2m/M. m 2␧ Moreover, the dependence v (E) in condensed rare gases d ͑ ͒ follow the indicated law ϳE→ϳͱE even more strictly, be- Already the threshold determined by condition 9 the con- ␭ ͑ ͒ dition ␣2Ӷ1 holds. This ͑as is seen from Eq. ͑3͒͒, is equiva- cause there is practically no dependence m(W) see Fig. 1 . The kinetic energy of the excess electron in steady drift lent to neglecting the energy lost by an electron in collisions increases in proportion to E, and the logical question arises: in comparison with the energy it acquires from the field, how high can it go? The statement first made in Ref. 12 is when essentially that this energy cannot appreciably exceed the w ͑ ϭ 0 1/2 ͑ ͒ first excitation potential I0 of the Rg atom for Xe, I0 Wi i 13 ϭ8.3 eV͒. As we see from Eq. ͑5͒, this should occur in fields & E that are very high, so that and 2 w0 2I ϵ␣Ϫ1у1. ͑9͒ Х 0 ϳ Ϫ2 ͑ ͒ ͱ␧ im 2 ; im E . 14 2I0 w0 In this case the excitation cross section increases so rap- Using relation ͑13͒ and changing the sums in ͑11͒ to inte- idly from the threshold that, upon exciting an atom, the elec- grals, we obtain 886 Low Temp. Phys. 27 (99–10), September–October 2001 E. B. Gordon and A. F. Shestakov

One of the first explanations for the existence of a lim- iting drift velocity of the electron in solid and liquid rare gases19 assumed the existence of structural traps that kept the electron trapped for a long time. This explanation, however, did not hold up to a critical scrutiny, since only impurities could serve as such traps; at the same time, as we have said, the presence of impurities only increases the drift velocity. Furthermore, that assumption does not account for the limi- tation of the drift velocity in gases. The model proposed in Ref. 12 and developed in the present paper in principle can explain the existence of rather FIG. 2. Electroluminescence spectra of gaseous Xe at pressures of 100 ͑1͒, deep traps in pure rare gases. Unlike the atoms in the ground 500 ͑2͒ and 760 torr ͑3͒.23 The arrows indicate the positions of the transition 3 →1 state, the electronically excited Rg atoms have a appreciable energies of the xenon atoms Xe( P1 S0), 8.44 eV, and of the negative Ϫ 2 → 1 ϩ electron affinity and, since immediately after excitation of an ions Xe ( P3/2) Xe( S0) e, 8.02 eV. atom the electron has practically no kinetic energy, it can be assumed that, at least in condensed rare gases, the electron can be trapped by the centers they themselves create. As an 3 w0 upper estimate of the lifetime of such a formation one can v Х ; v ϳE. ͑15͒ d 25/2 ͱ d ␶ Ϫ mI0 take the radiation lifetime R of the negative ion (Rg2*) .If it is assumed that this time is close to that for the two-center In other words, contrary to crass intuition, when an upper Ϫ exciton ͑␶ Х(2–3)ϫ10 8 s for Xe*͒, then we find that, limit is placed on the random motion of the electron its drift R 2 starting at an electric field Eϭ103 V/cm, the lifetime of a velocity does not remain constant but increases in proportion trapped electron in liquid and solid Xe become comparable to the applied field, i.e., faster than for stationary drift. The to the time it takes for a drifting electron to acquire an energy yield of excimer molecules ͑two-center excitons͒ and, ac- equal to the energy of electronic excitation of the xenon cordingly, the number of VUV photons per drifting electron, atom. starting from their appearance threshold ͑9͒, should also It is not hard to obtain an expression for the drift veloc- grow in proportion to the field: ity of electrons with allowance for this effect: eEl ␩ϭ ͑ ͒ ͚im ϩ␶ ␶ , 16 1 iϭ0ti R 1 R 1 eE I0 ϭ ϭ ϩ Х ϩ . ͑17͒ im 0 im 0 vd ͚ s v ͚ s v I0 where l is the thickness of the sample. iϭ0 i d iϭ0 i d Let us compare the predictions of our simple proposed We see from Eq. ͑17͒ that taking into account the effect of model with the results of experiments. As we see in Fig. 1, electron trapping by the excited xenon ͑first by the atom, and ͒ where as an example we present the experimental curves of then by the Xe2* molecule gives a term that causes a de- ͑ ϳ Ϫ1 ͒ vd(E) for gaseous, liquid, and solid xenon, the linear growth crease in the limiting drift velocity vd E at large E . segment ͑corresponding to isothermal drift͒ does indeed give Without assuming a substantial dependence of the radiation Ϫ ͑ way to a square-root dependence on E. Moreover, in gaseous lifetime of (Xe2*) on E which is unlikely for an allowed xenon at a comparatively low pressure, when the drift veloc- optical transition͒ it is hard to explain the observed con- ity reaches a value approximately equal to that found from stancy of the drift velocity at high electric fields ͑for Xe, in ͑ ͒ ϭ ϭͱ Ϸ ϫ 5 3 4 2 Eq. 8 for W I0 , i.e., vd I0 /M 5 10 cm/s, the func- the range 1ϫ10 –6ϫ10 V/cm ͒ by a competition of the Ϫ1 tion vd(E) actually does again become linear, and it is in this contributions proportional to E and E . For gaseous Xe the range of electric field E that the VUV radiation in the 170 formation of long-lived negative ions is unlikely altogether. ͑ ͒ nm region the luminescence of Xe2* appears in a threshold This leaves us to consider the problem of the interaction manner, its yield depending linearly on the applied voltage; of Rg atoms with an electron in more detail. The metastable and, in agreement with the above model, this dependence is states of the corresponding negative ions are well known, 2 the same when the gas pressure is increased at least to 1.4 and the lowest in energy of them are the resonance P3/2 bar.20 Furthermore, at low pressures one observes a spectral states with the configuration np5(nϩ1)s2 ͑an additional band in the 150 nm region, i.e., in the region where the electron is found on the same orbit as an outer electron of the 6 5 2 3 ͒ ⌬ atomic transition Xe(5p –5p ( P3/2)6s) is found, shaded to lowest P2 state of the excited Rg atom lie an amount 20 ͑ ͒ ϭ 3 the red side see Fig. 2 . 0.4 eV lower in energy than the excited P2 Rydberg state However, in gaseous xenon at high pressure, as in con- of the atom ͑see Table I͒. They have a rather long lifetime23 densed xenon, the vd(E) curve goes over not to a linear and decay as a result of the weak Coulomb interaction of the trend as implied by the proposed model but rather to a con- electrons in a process involving the removal of one of the stant value. This behavior is even more strange in that it electrons and a transition of the other to the ground state. In contradicts simple physical ideas—an increase of the energy two-particle collisions these resonances are manifested as loss, by decreasing the effective number of random walks, rather sharp ͑with a width of about 0.01 eV͒ and therefore should always lead to an increase in the drift velocity. This is intense peaks in the elastic and inelastic electron scattering ␴ also confirmed by the results of experiments—the introduc- cross sections. These scattering cross sections, R tion of a molecular impurity in rare gases always leads to an ϭ10Ϫ15 cmϪ2, are much larger than the cross sections typi- ͑ ͒ 17 increase in vd Ref. 21 . cal for the region after the Ramsauer minimum. Neverthe- Low Temp. Phys. 27 (99–10), September–October 2001 E. B. Gordon and A. F. Shestakov 887 less, in gaseous Xe at densities up to 1020 cmϪ3, on account ϩ → *ϩ e*¯Xe Xe Xe2 e of the energy acquired by the electron from the field over the ͑since the energy of Xe* at 0.7 eV ͑see Table I͒ is lower than mean free path in fields of the order of 1 kV/cm, i.e., W 2 Ϫ the energy of atomic excitation, the final energy of the elec- ХE/␴ nϭ10 2 eV, there is a high probability of ‘‘slipping R tron is 0.3 eV, which is quite sufficient for delocalization of past’’ these so-called core excited resonances. Accordingly, the latter͒. in this case the model presented above is valid, and the ex- The published data available24 on the transformation of perimental v (E) curves have a regular form ͑see Fig. 1͒, the d the VUV emission spectrum generated by electrons drifting transition to a linear dependence of v on E occurring pre- d in gaseous xenon as the Xe density is increased agree sur- cisely in the region where the VUV emission due to the prisingly well with the arguments presented above. Indeed, electronic excitation of xenon appears. as can be seen in Fig. 2, at low densities the electrolumines- In a denser gas and in the condensed state, the motion of cence is dominated by a narrow band with ␭ϭ150 nm, cor- an excess electron in the energy region corresponding to the responding to the transition of the excited atom. When the resonant scattering peak in the configuration of the negative density is increased on account of a weakening of this band ion can be regarded as a sharp transition to its drift with a ͑the total radiation yield is independent of the density͒ there markedly smaller step of the random walk. As we see from arise: 1͒ a wide band, close in shape and position (␭ Eq. ͑3͒ for the electron energy, the first term on the right- ϭ170 nm) to the emission spectrum of the excimer molecule hand side of the part corresponding to the acquisition of Ϫ Xe* , and 2͒ a narrower peak centered at ␭ϭ160 nm, which energy from the field decreases as ␴ 2 , while the second 2 R corresponds precisely to the energy of a state of the negative term, describing the loss of energy, decreases more weakly Ϫ Ϫ ion (Xe*) . Finally, at high densities there remains only the than ␴ 1 . ͑The electron energy loss in the region of the R band corresponding to the emission of the excimer molecule. resonances due to states of the negative ion is small—for the In the condensed state, as in the dense gas, the main decay of a negative ion the momentum conservation law channels of electron energy loss are the formation and sub- gives, as before, ␧ϳm/M.͒ This means that during drift in sequent decay of a one-center negative ion (Xe*)Ϫ and a the resonance region the electron energy tends to decrease, two-center exciton. This is confirmed by the spectrum of thereby removing it from the resonance region. As a result of quasi-molecular emission observed when solid Xe is irradi- this negative feedback, the electron energy W remains con- ated by fast electrons.1 With allowance for the matrix shift, stant, WϭI Ϫ⌬, over a rather long time, regardless of the 0 the L band corresponds well to the position of the transition value of the electric field. Accordingly, the average drift ve- x in (Xe*)Ϫ ͑see Fig. 3͒. However, in the liquid and solid the locity v of the electron after this time reaches its steady d electron can with a certain probability be localized near a value given by formula ͑7͒ with a certain effective value of binuclear center. As we have said, the excited state (Xe*)Ϫ the parameter ␧. Thus, taking into account the scattering on 2 rapidly decays: structure resonances of the negative ion indeed leads to con- ͑ ͒Ϫ→ ϩ stancy of vd at large E, and only when the field is increased Xe2* 2Xe e, by a factor ␤ϭ␴ /␴ ͑␴ and ␴ are the cross sections for R 0 0 R leading to a radiationless energy loss of around 0.4 eV and ordinary and resonance scattering, respectively͒ does the thereby reducing the yield of excitons per electron. drift velocity of the electron begin to grow as ͱE. In dense gases and in the condensed state there is appar- During drift motion in a dense gas, with a kinetic energy ently one more important cause of the decrease in the effi- remaining constant at WϭI Ϫ⌬ for a long time, an electron 0 ciency ␩ of energy conversion of electric field into VUV can either: 1͒ skip the resonance peak because of fluctuations radiation as the electric field E is increased. In the steady of its energy, and having acquired insufficient energy to ex- drift of an electron near WϭI Ϫ⌬ the distance traveled by it cite an rare-gas atom ͑and this, as we have said, subsequently 0 in the direction of the field does not, of course, lead to a leads to emission at a transition of the atom or excimer mol- change of the electron energy. Therefore, in high fields the ecule͒;2͒ having been incorporated into a metastable nega- value of ␩ is determined by the ratio of the distance traveled tive ion, undergo a transition to a free state in the process of its decay, with the emission of a VUV photon; 3͒ with a constant velocity of around 10Ϫ30 cm/s enter into a trimo- lecular reaction forming an excimer molecule:

TABLE I. Characteristics of the excited states of atoms of the rare gases Rg in eV1.

Energy of lowest 3 ( P2) Energy of Energy of excited states dissociation of phototransition Electron of the the exciplex in the affinity of ͑ ͒ Rg atom Rg* Rg2*,D0 exciplex Rg* Ref. 22 Ar 11.624 1.2 9.76 0.44 Kr 10.032 0.9 8.38 0.40 FIG. 3. Quasi-molecular emission of crystalline Xe at 10 K ͑1͒ and76K ͑ ͒ 1 Xe 8.440 0.7 7.2 0.455 2 . The ‘‘hot’’ band Lx can be interpreted as emission at a transition of the negative ion (Xe*)Ϫ. 888 Low Temp. Phys. 27 (99–10), September–October 2001 E. B. Gordon and A. F. Shestakov by the electron in the period of its intense absorption of energy from the field, to the total distance, the main contri- bution to which is from the aforementioned steady drift stage:

͚im iϭ0si 1 ␩ϭ ϭ , ͑18͒ ͚im ϩ 0␶ 1ϩ͑v0␶eE͒/I iϭ0si vd d 0 where ␶ is the time prior to trapping of the electron. Accord- ingly, the number of photons, m, generated by one electron drifting through a sample of thickness l tends, in high fields, toward a constant value

l mϭ . ͑19͒ FIG. 4. Dependence of the threshold electric field for the appearance of 0␶ VUV emission as a function of the density n of Xe at Tϭ290 K. The arrow vd indicates the position of the critical density nc . It is easily estimated that mϷ100, which in principle would permit obtaining an electron avalanche in a cell with a pho- increases in proportion to the density in accordance with tocathode, as was proposed in Ref. 12. Refs. 20 and 26, at densities of 5ϫ1021 cmϪ3 it has already The analysis in this paper does not leave any doubt that become 3 orders of magnitude higher, reaching values of р 20 Ϫ3 in gaseous xenon at densities n 10 cm the characteris- 1000 kV/cm, which would be hard to achieve experimen- tics of the drift of an excess electron in a high electric field is tally. However, once the liquid has formed, the threshold determined by the electronic excitation of the gas, which is returns to a value of around 1 kV/cm ͑in agreement with Ref. thus an efficient converter of energy from a static electric 14͒. In other words, in compressed rare gases the drift of field directly into VUV radiation. The conclusion that a simi- electrons is unaccompanied by luminescence all the way up lar mechanism is realized in liquid and solid xenon is also until condensation. The calculations whose results are pre- quite convincing, particularly since it can explain a whole set sented in Fig. 4 were done for a temperature close to the of existing experimental results. Among the most important ϭ critical temperature Tc 290 K. The analysis of the behavior of these is the recently published direct observation of VUV of Eth(n) in the condensed state should be done at tempera- emission in the two-center excitonic band in liquid xenon tures Tϭ160– 200 K, close to those used in the experiments 14 during the drift in it of electrons emitted by a cold cathode. on the measurement of the mobility.11 It turns out that, al- At the present time it is not possible to estimate the field though in the liquid and solid states the value of Eth depends corresponding to the threshold observed in Ref. 14 for the nonmonotonically on n, it remains low—around 1 kV/cm. appearance of VUV electroluminescence, since a more de- We have based the above analysis on experimental data 25 tailed acquaintance with the experimental technique used obtained for the drift of electrons in xenon, the most studied by the authors of Ref. 14 indicates the possibility of consid- of the heavy rare gases. However, the basic conclusions are erable nonuniformity of the electric field, although the also valid for Kr and Ar, which are potential sources of even threshold cannot be very high, since the voltage used was harder UV radiation. only 70 V. If this line of research is successful, experimenters will Nevertheless, the final answer to the question will re- acquire a new way of studying electronic excitations in liq- quire doing experiments with solid or liquid Xe in a uniform uid and solid rare gases, the features of which include com- field with a determination of the absolute value of the VUV parative simplicity of the experiment and high intensity of emission yield. We are currently doing such experiments in emission on the one hand, and a specific means of excitation, conjunction with the Kamerlingh Onnes Leiden University. including the physical presence of an electron at the site of Let us conclude by noting an interesting feature of the localization of the exciton, on the other. The high efficiency VUV electroluminescence in Xe, which follows from the of excitonic luminescence not only facilitates the recording ͑ ͒ ͑ ͒ above analysis. By combining formulas 4 and 9 , one can of its characteristics but holds promise for a number of prac- obtain the following expression for the threshold value of the tical applications, such as the creation of excitonic VUV la- field at which this emission appears: sers and lamps and the refinement of luminescent detectors of high-energy particles. v I ͱ␧ ϭ T 0 ͑ ͒ The authors thank P. Wyder ͑Grenoble High Magnetic Eth , 20 2kT ␮ Field Laboratory͒ for providing the dissertation of A. S. Schussler, which contains the details of their work. This ␮ϵ where vd /E is the electron mobility measured in a low study was supported in part by the Russian Foundation for ͑ ͒ field. Figure 4 shows a plot of Eth(n) according to Eq. 20 Basic Research, grants Nos. 01-03-32247 and 00-15-97400. with the use of the experimental values of ␮(n) from Ref. 17, where n is the density of Xe. We see that whereas at low *E-mail: [email protected] Ϫ densities ͑1019 cm 3͒ the threshold is around 1 kV/cm and **E-mail: [email protected] Low Temp. Phys. 27 (99–10), September–October 2001 E. B. Gordon and A. F. Shestakov 889

1 E. V. Savchenko, I. Ya. Fugol’, in Cryocrystals edited by B. I. Verkin and 14 A. S. Schussler, J. Burghorn, P. Wyder, B. I. Lembrikov, and R. Baptist, A. F. Prikhod’ko ͓in Russian͔, Naukova Dumka, Kiev ͑1983͒, Part 4. Appl. Phys. Lett. 77,2786͑2000͒. 2 ͑ ͒ 15 ͓ ͔ Ya. Frenkel, Phys. Rev. 37,1279 1931 . Yu.P.RaŽzer, Physics of the Gas Discharge in Russian , Nauka, Moscow 3 N. B. Basov, V. A. Danilychev, Yu. M. Popov, and D. D. Khodkevich, ͑1992͒. JETP Lett. 12,329͑1970͒. 16 B. M. Smirnov, Physics of Weakly Ionized Gases ͓Mir, Moscow ͑1981͒; 4 H. Nahme and N. Schwentner, Appl. Phys. B: Photophys. Laser Chem. 51, Nauka, Moscow ͑1978͔͒. 177 ͑1990͒. 17 S. S.-S. Huang and G. R. Freeman, J. Chem. Phys. 68, 1355 ͑1978͒. 5 N. Schwentner and V. A. Apkarian, Chem. Phys. Lett. 154, 413 ͑1989͒. 18 S. S.-S. Huang and G. R. Freeman, Phys. Rev. A 24, 714 ͑1981͒. 6 ͓ 19 ͑ ͒ A. V. EletskiŽ and B. M. Smirnov, Physical Processes in Gas Lasers in J. Lekner, Phys. Rev. 158, 130 1967 . Russian͔,E´ nergoatomizdat, Moscow ͑1985͒, p. 124. 20 F. Favata, A. Smith, M. Bavadaz, and T. Z. Kowalski, Nucl. Instrum. 7 C. K. Rhodes ͑Ed.͒, Excimer Lasers, Springer-Verlag, Berlin ͑1979͒. Methods 294, 595 ͑1990͒. 8 J. P. Hernandez, Rev. Mod. Phys. 63, 675 ͑1991͒. 21 D. W. Swan, Proc. Phys. Soc. ͑London͒ 83, 659 ͑1964͒. 9 G. Ascarelli, J. Phys. Chem. 84,1193͑1980͒. 22 H. S. W. Massey, Negative Ions, Cambridge University Press, Cambridge 10 J. Bardeen and W. Shockley, Phys. Rev. 80,72͑1950͒. ͑1976͒, Mir, Moscow ͑1979͒. 11 ´ 23 ͑ ͒ E. M. Gushchin, A. A. Kruglov, and I. M. ObodovskiŽ, Zh. Eksp. Teor. E. B. Gordon and O. S. Rzhevsky, Molecular Physics 2001 . Fiz., 82, 1114 ͑1982͓͒Sov. Phys. JETP 55, 650 ͑1982͔͒. 24 B. A. Monich, Prib. Tekh. E´ ksp. 5,7͑1980͒. 12 E. B. Gordon, V. V. Khmelenko, and O. S. Rzhevsky, Chem. Phys. Lett. 25 A. S. Schussler, Thesis, Hartung-Gorre Verlag, Konstanz ͑1998͒. 217,605͑1994͒. 26 A. El-Habachi and K. H. Schoenbach, Appl. Phys. Lett. 73,885͑1998͒. 13 E. B. Gordon, O. S. Rzhevsky, and V. V. Khmelenko, Kvantovaya E´ lek- tron. ͑Moscow͒, 21,227͑1994͒. Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Ionization and hydrolysis of dinitrogen pentoxide in low-temperature solids J. Agreiter, M. Frankowski,* and V. E. Bondybey**

Institute for Physical and Theoretical Chemistry, Technical University of Munich, Lichtenbergstrasse 4, 85747 Garching, Germany ͑Submitted June 21, 2001͒ Fiz. Nizk. Temp. 27, 1202–1207 ͑September–October 2001͒ Solid layers of interest for the chemistry of polar stratospheric clouds are investigated. Mixtures of covalent N2O5 and water in various molar ratios are deposited from the vapor phase on a cold 12 K substrate. By repeatedly recording Fourier transform infrared spectra of the samples during gradual warmup to 200 K over a period of several hours the hydrolysis process can be followed. At each concentration the process is found to proceed in several distinct steps with sharp temperature thresholds. In samples containing only small amounts of water the covalent N2O5 is first, at around 110 K, converted to an ionic nitronium nitrate, and only in a ϩ Ϫ subsequent step does NO2 NO3 react to form nitric acid. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1414583͔

INTRODUCTION the so-called NOx or ‘‘odd nitrogen’’ species. These, and in The structures, optical spectra, and other properties of particular NO2, in turn have importance in reacting with ClO molecules in condensed phases often differ quite substan- to form the relatively unreactive chlorine nitrate, ClONO2, tially from those in the gas phase. Strongly polar compounds thus removing reactive chlorine from a catalytic, ozone de- like HCl or Hl, which occur as covalent compounds in the stroying cycle. gas phase, ionize spontaneously when dissolved in water. Perhaps somewhat paradoxically for the rarefied strato- Even though the autoionizing reaction of water itself, sphere, the properties of N2O5 and other nitrogen oxide in → ϩϩ Ϫ ϭ the solid phase or on solid surfaces may be even more im- 2H2O H3O OH requires an energy H 933 kJ/mol, ev- portant for stratospheric chemistry than their gas-phase ery liter of pure room-temperature water contains in thermal 3–8 17 ϩ reactions. When in the course of the polar winter the tem- equilibrium some 10 H3O cations and a corresponding number of OHϪ anions, which greatly affect its properties. perature of the stratosphere drops below about 190 K, the ͑ ͒ Some substances, which in the gas phase are present in a formation of so-called polar stratospheric clouds PSCs covalent form tend to ionize completely in a solid or liquid, takes place. These consist of micron-size particles, whose where the substance itself acts as its own ‘‘solvent.’’ While in main components are water and the oxides of nitrogen. Het- the gas phase a covalent form AB is usually lower in energy, erogeneous processes on their surfaces are then believed to in a condensed phase where an Aϩ ion can interact with a be responsible for converting relatively unreactive, so-called number of nearby BϪ anions, the lattice energy may more ‘‘reservoir’’ chlorine containing compounds, mainly chlorine than compensate for this deficit, and the ionic AϩBϪ alterna- nitrate and hydrogen chloride, into more active species, such tive becomes energetically preferable. A well-known ex- as Cl2 or NOCl. During polar spring, when the temperature rises again and the PSCs evaporate, chlorine in this more ample of this type is dinitrogen pentoxide, N2O5, which oc- reactive form is returned into the gas phase, with photolysis curs in the gas phase in the form of covalent O2N–O–NO2 molecules, but in the solid usually in the ionic form yielding atomic chlorine. Cl atoms then very efficiently cata- ϩ Ϫ NO NO , nitronium nitrate.1,2 lyze O3 destruction and are ultimately responsible for the 2 3 seasonal ozone depletion and catalytic destruction, referred to popularly as the ‘‘ozone hole.’’ NITROGEN OXIDE IN THE STRATOSPHERE Also the relatively nonvolatile dinitrogen pentoxide con- Nitrogen oxides and their condensation are of consider- denses on the surface of PSCs and contributes in a variety of able importance in the chemistry of the stratosphere and con- ways to the ozone and its balance. Important is the hydroly- tribute in several ways to the balance of stratospheric sis of N2O5 to form nitric acid, presumably by the reaction 3–8 11–13 ozone. The major path of N2O5 formation is due to reac- catalyzed on a PSC surface: tion of nitric oxide with ozone:9,10 N O ϩH O→2HNO . ϩ → ϩ 2 5 2 3 NO2 O3 NO3 O2, This reaction is in fact believed to be the major pathway NO ϩNO ͑ϩM ͒→N O ͑ϩM ͒. 2 3 2 5 of ‘‘denitrification’’ of the stratosphere, that is, for the loss of Since during the daytime NO3 is efficiently photolyzed NOx species, and thus to contribute indirectly to an increase by sunlight, the N2O5 concentration reaches its maximum at in the concentrations of active chlorine radicals and to ozone night. The dinitrogen pentoxide is ultimately also photolyzed destruction. The nitric acid concentration also affects the ϩ back to NO2 NO3, but it acts as a relatively unreactive PSC formation itself: the so-called type-II PSCs, containing ‘‘reservoir’’ reducing, albeit temporarily, the concentration of nitric oxides and, in particular, nitric acid trihydrate ͑NAT͒,

1063-777X/2001/27(9–10)/5/$20.00890 © 2001 American Institute of Physics Low Temp. Phys. 27 (99–10), September–October 2001 Agreiter et al. 891 can form at a higher temperature than ‘‘pure water,’’ type-I PSCs. While the above hydrolytic reaction forming nitric acid could in principle also occur during gas phase colli- sions, this is believed to be far too inefficient to explain the observed chemistry. This added practical importance of N2O5 and of its con- densation and hydrolysis has motivated a number of recent experimental and theoretical studies. Thin layers consisting of nitrogen oxides, nitric acid and water were studied by spectroscopy, and their reactions were also extensively inves- tigated at a variety of temperatures. Also theoretical, ab initio ͑ ͒ or density functional theory DFT studies of N2O5 and its hydrolysis have appeared.14 In spite of this extensive atten- tion, numerous open or controversial problems remain. A considerable discussion centers on the question of whether the reactants occur and the reaction products are formed in FIG. 1. Experiment with H2O:N2O5 ratio of about 0.25:1. The initial 10 K infrared spectrum of covalent N2O5 in the bottom trace is in the top trace an ionized or in a molecular form. Based on an infrared ϩ Ϫ after gradual annealing over4hto160Kconverted to ionic N O NO , study it was concluded that N O itself does not play a major 2 3 2 5 with small amounts of nitric acid, HNO3. role in the heterogeneous chemistry directly but rather through its hydrolysis products. The structure and reactivity of nitrogen oxides is known to be a sensitive function of their and displayed in a variety of ways. Initial spectra, final spec- temperature and of the dynamics of their formation. To gain tra, or intermediate spectra at any time during the warmup additional insights, we reinvestigate here the N2O5 hydroly- could be plotted. Alternatively, by plotting the absorption at a sis using Fourier transform infrared ͑FTIR͒ spectroscopy. selected wavelength, the decay of a reacting compound or ϩ Ϫ Even though in solid the ionic NO2 NO3 form is lower the appearance a product as a function of time and tempera- in energy, if gaseous dinitrogen pentoxide is condensed on a ture could be displayed. Finally one could produce a ‘‘three- very cold surface, a solid consisting of discrete, covalent dimensional’’ diagram with the abscissa indicating the wave ͑ N2O5 molecules is formed, whose infrared or Raman spectra number, the ordinate the time or the temperature, which was are very similar to the gas-phase spectrum. When the energy rising approximately linearly with time͒ and the ‘‘color’’ or needed to overcome the activation barrier is supplied, either shade of gray the band intensity. In the samples studied here by optical irradiation or by heating the sample, it is con- there was little change in the spectrum up to about 100 K. To verted into the ionic form. As is well known, annealing such shorten the duration of the experiments, the sample tempera- samples above about 120 K results in an abrupt and complete ture was first ‘‘rapidly,’’ over about ten minutes, raised to 80 change in the structure of the solid, and naturally also in its K. After that it was allowed to rise further, over the course of infrared or Raman spectrum. In the present work we take about 4–6 h, to about 160–200 K, at which point the experi- advantage of the ability of a modern FTIR instrument to ment was terminated and the sample allowed to evaporate. rapidly acquire and store digitally infrared spectra. We pre- pare solid mixtures of water and covalent N2O5 at about 12 RESULTS AND DISCUSSION K by deposition from the gas phase. Then we slowly allow the samples to warm up over several hours, and we repeti- Figure 1 exemplifies the results of such an experiment tively take infrared spectra every few minutes. The spectra obtained with a sample containing N2O5 with a relatively ͑ ͒ then provide information about structural changes, phase low water concentration H2O:N2O5 molar ratio 0.25:1 de- ϳ transitions, or chemical reactions taking place in the solid. posited at a temperature of 12 K. The initial spectrum in The interpretation of the results is naturally greatly assisted the bottom trace of Fig. 1 shows the known bands of the Ϫ1 by the availability of numerous previous studies of the spec- covalent N2O5, e.g., near 1740, 1240, and 756 cm , most of troscopy of nitrogen oxides, nitric acid, and of its hydrates. which are only slightly shifted with respect to the gas phase. The presence of water is evidenced by the broad band cen- tered near 3350 cmϪ1. EXPERIMENTAL As the sample is warmed up, the spectrum initially re- The N2O5, prepared in a standard way by the reaction of mains unchanged until a temperature of about 110 K is ͑ ͒ commercial nitric oxide or N2O4 with ozone, was purified reached. Here over a temperature range of less than 10 K the by vacuum distillation and by several pump–freeze–thaw covalent N2O5 bands almost completely disappear and are cycles. The oxide stored in Pyrex containers at liquid nitro- replaced by absorptions of the ionic solid, the most charac- gen temperature was then deposited via a corrosion-resistant teristic being the asymmetric stretch of the nitronium cation ϳ ϩ Ϫ1 needle valve onto a metal substrate cooled to 12Kbya NO2 at 2386 cm and the v2 vibration of the nitrate anion Ϫ1 closed-cycle refrigerator. Simultaneously with the N2O5, wa- near 820 cm . A simple program was written which dis- ter was deposited from a separate inlet, to yield a solid mix- plays the information obtained from the repetitive spectral ture in a desired molar ratio of H2O:N2O5. Infrared spectra scans during the sample warmup in the form of a colored two were recorded on a Bruker IFS 120 HR Fourier transform dimensional diagram, in which the spectral changes occur- spectrometer. The 100–300 spectra acquired during over ring in the sample can be very nicely seen. As an example we 4–6 h during the slow sample warmup could then be used present in Fig. 2 a black and white version of one of these 892 Low Temp. Phys. 27 (99–10), September–October 2001 Agreiter et al.

FIG. 3. Intensity profiles showing the simultaneous growth of the ϩ Ϫ ϳ N2O NO3 bands within a relatively narrow, 5 K temperature range.

of about 3:1, stoichiometry which should lead to nitric acid monohydrate ͑NAM͒: FIG. 2. ‘‘3D’’ diagram, from the same experiment showing the changes of ϩ ϩ → ϩ Ϫ N2O5 3H2O 2HNO3•H2O 2H3O NO3 . spectrum during the controlled warm-up. The initial bands of covalent N2O5 disappear after ϳ120 min ͑at 110 K͒, with concurrent appearance of the Such a sample when deposited at 12 K shows again a ionic nitronium nitrate. After about 220 min ͑ϳ140 K͒ nitric acid bands superposition of the spectrum of covalent N O , with in this appear. 2 5 case a much more prominent water band near 3360 cmϪ1,as shown in Fig. 4. Interestingly, a minor amount of the ionic form seems to be present, perhaps small ‘‘clusters’’ whose diagrams. Even though in the shades of gray version the conversion during deposition is ‘‘catalyzed’’ by the water- changes are much less apparent, one can still follow the dis- rich surface. During the warmup little happens until about appearance of the covalent solid and growth of the ionic 130 K is reached, when major changes in the spectrum begin solid as a function of time. to take place. In the first step the bands of the covalent N2O5 With some time delay after the conversion to ionic N2O5 start to disappear, with simultaneous appearance of the ab- is complete, a second, less prominent change occurs in the sorptions of nitric acid. The concentration of HNO3, which spectrum near 140 K, when bands of nitric acid begin to can be followed by monitoring its absorptions near 775 and grow in. While some of these are partially overlapping, the 945 cmϪ1, reaches a maximum near 145 K and then starts to HNO3 absorptions are easily identified near 775 and 945 disappear again. Before the final temperature of 180 K is Ϫ1 cm . Another strong HNO3 band, nearly coincident with an reached, the nitric acid is completely replaced by NAM. This N O absorption, occurs around 1680 cmϪ1. The final spec- ϩ 2 5 is best identified by the H3O absorptions near 1670, 2250, trum resulting from annealing of the sample to 160 K can be and 2650 cmϪ1 and the nitrate anion bands near 735, 812, seen in the top trace of Fig. 1, where the nitric acid bands can and 1260 cmϪ1. The absorptions of NAM then remain until be easily identified. These observations provide clear evi- the sample is lost. dence that under the conditions of our experiments, the neu- tral dinitrogen pentoxide does not react with the water present in the sample. The formation of nitric acid only takes place with a considerable delay, after a complete conversion of the nitrogen pentoxide solid into an ionic form. A different presentation of these data and of the conver- sion of the covalent dinitrogen pentoxide into the ionic ni- tronium nitrate is shown in Fig. 3. Here one can see the simultaneous and abrupt appearance of the ionic solid ab- sorptions at 820, 1400, and 2386 cmϪ1 after about 115 min during controlled sample heating. The spectral change occurs within a narrow range around 110 K, and concurrently one can observe a decrease in the absorptions due to the covalent pentoxide. Note that the thin solid line gives the changes in the sample temperature, as indicated by the right-hand scale. The formation of nitric acid hydrates which are particu- larly important in the formation and chemistry of polar stratospheric clouds, and apparently in the ozone balance, FIG. 4. Experiment with H2O:N2O5 ratio of 3:1, corresponding to nitric acid monohydrate ͑NAM͒. The initial, bottom spectrum shows mainly covalent can be studied by increasing the proportion of water in the Ϫ N O , and a very strong, broad water band near 3360 cm 1. The middle deposited H O:N O samples.15–18 Interesting are the 2 5 2 2 5 spectrum after 330 min exhibits maximum HNO3 concentration. The top intermediate-concentration samples with an H2O:N2O5 ratio spectrum after 400 min is mainly that of the NAM. Low Temp. Phys. 27 (99–10), September–October 2001 Agreiter et al. 893

FIG. 6. Time resolved intensity profiles showing the NAD absorptions growing in after ϳ280 min, reaching a maximum after ϳ305 min ͑ϳ145 K͒ and then abruptly disappear with concurrent appearance of the NAT. FIG. 5. H2O:N2O5 ratio of about 8, with enough water to form nitric acid trihydrate, NAT. The initial, bottom spectrum contains covalent N2O5, and a strong water band. In the middle spectrum after about 300 min ͑ϳ140 K͒ the band, at a temperature near 145 K the NAD appears but only dinitrogen pentoxide has disappeared, and the observed absorptions agree Ϫ1 with the known spectrum of NAD. After 320 min ͑ϳ145 K͒ another sharp transiently, only to be replaced by the strong 1100 cm band change takes place and the top spectrum identified as NAT results as final of the final product, NAT. product. SUMMARY In summary, in this study we have investigated the struc- A considerably different behavior can be found in tural changes in solid layers with relevance to the chemistry samples containing still larger relative concentrations of wa- of PSCs. We condensed the samples at low temperatures and ter, where one can naturally observe a higher degree of nitric then repeatedly recorded their infrared spectra as a function acid hydration. As in the less concentrated samples, the re- of time as the sample temperature was slowly raised, to gain action again proceeds in several distinct steps. Such an ex- insight into processes and structural changes that take place. periment is presented in Fig. 5, where the molar ratio of H2O We found that the hydrolysis of N2O5 proceeds in each case to N2O5 was about 8:1, allowing the formation of nitric acid in several well-defined sequential steps. Specifically, in trihydrate: samples of dinitrogen pentoxide there is a distinct transition ϩ Ϫ to an ionic NO2 NO3 form around 120 K. When only small ϩ → → ϩ Ϫ amounts of water ͑ϳ0.2:1͒ are present, a further reaction to N2O5 7H2O 2HNO3 3H3O 2H7O3 NO3 . • form nitric acid occurs only with delay, after a complete conversion into the ionic form. When enough water is In this case conversion of N O from covalent to ionic 2 5 present ͑Ͼ3:1͒ the nitric acid produced in the first step after form is hardly observed at all, with the covalent form per- some delay reacts further with water to form NAM. In sisting up to about 150 K. Apparently under these conditions samples containing a large excess of water ͑Ͼ7:1͒ almost no one effectively has discrete N O monomers ‘‘isolated’’in 2 5 conversion of the pentoxide into an ionic form takes place. water glass. As the temperature is gradually raised, the The covalent form persists up to about 150 K, where it reacts changes start only above 140 K, but unlike in the more con- directly with H O first to yield nitric acid dihydrate and, in a centrated N O case, neither nitric acid nor nitric acid mono- 2 2 5 second reaction step, finally to trihydrate. hydrate is detected. The new bands, which grow in, can on 15–19 the basis of previous spectroscopic work be easily iden- * ͑ ͒ Permanent address: Institute of Fluid-Flow Machinery, Polish Academy of tified as nitric acid dihydrate NAD . The most characteristic Sciences, J. Fiszera 14, 80-231 Gdan´sk, Poland feature is the presence of two strong bands near 1290 and **E-mail: [email protected] 1420 cmϪ1, interpreted previously in terms of the asymmet- ric nitrate stretching mode split by the asymmetric environ- ment. However, the NAD bands do not persist long but are, 1 R. Teranishi and J. C. Decius, J. Chem. Phys. 22, 896 ͑1954͒. at 155 K, abruptly replaced by the spectrum of the final 2 E. Grison, K. Eriks, and J. L. de Vries, Acta Crystallogr. 3,290͑1950͒. 3 product, nitric acid trihydrate. In NAT the two nitrate stretch- S. Solomon, R. R. Garcia, F. S. Rowland, and D. J. Wuebbles, Nature ͑London͒ 321, 755 ͑1986͒. ing bands collapse to a single strong absorption at 1390 4 M. Molina, T. L. Tsao, L. T. Molina, and F. C. Y. Wang, Science 238,1253 Ϫ1 cm , and two bands appear in the OH stretching region, a ͑1987͒. sharper one at 3424 cmϪ1 and a much broader band near 5 S. Solomon, Nature ͑London͒ 347, 347 ͑1990͒. Ϫ 6 ͑ ͒ 3205 cm 1. The apparently quite stable trihydrate, NAT, M. A. Tolbert, Science 264, 527 1994 . 7 A. G. Russell, G. J. McRae, and G. R. Cass, Atmos. Environ. 19, 893 which is believed to be one of the major components of the ͑1985͒. polar stratospheric clouds, then remains unchanged until the 8 T. F. Mentel, D. Bleilebens, and A. Wahner, Atmos. Environ. 30,4007 sample is lost. ͑1996͒. 9 Similarly to Fig. 3, Fig. 6 shows the time dependent C. R. Webster, R. D. May, R. Toumi, and R. Pyle, J. Geophys. Res. 95, 13851 ͑1990͒. change in the product infrared absorption intensities for the 10 F. J. Dentener and P. J. Crutzen, J. Geophys. Res. 98, 7949 ͑1993͒. Ϫ 8:1 sample of Fig. 5. As demonstrated by the 1030 cm 1 11 M. T. Leu, Geophys. Res. Lett. 15, 851 ͑1988͒. 894 Low Temp. Phys. 27 (99–10), September–October 2001 Agreiter et al.

12 D. R. Hanson and A. R. Ravishankara, J. Geophys. Res. 96, 5081 17 N. Barton, B. Rowland, and J. P. , J. Phys. Chem. 97,5848͑1993͒. ͑1991͒. 18 R. H. Smith, M. T. Leu, and L. F. Keyser, J. Phys. Chem. 95,5924 13 M. A. Quinlan, C. M. Reihs, D. M. Golden, and M. A. Tolbert, J. Phys. ͑1991͒. Chem. 94, 3255 ͑1990͒. 19 A. B. Horn, T. Koch, M. A. Chesters, M. R. S. McCoustra, and J. R. 14 J. P. McNamara and I. H. Hillier, J. Phys. Chem. 104, 5307 ͑2000͒. Sodeau, J. Phys. Chem. 98, 946 ͑1994͒. 15 G. Ritzhaupt and J. P. Devlin, J. Phys. Chem. 95,90͑1991͒. 16 B. G. Koehler, A. M. Middlebrook, and M. A. Tolbert, J. Geophys. Res. This article was published in English in the original Russian journal. Repro- 97,8065͑1992͒. duced here with stylistic changes by AIP. LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Weak ferromagnetism in the antiferromagnetic magnetoelectric crystal LiCoPO4 N. F. Kharchenko and Yu. N. Kharchenko*

B. Verkin Institute for Low Temperature Physics and Engineering, National Academy of Sciences of Ukraine, pr. Lenina 47, 61103 Kharkov, Ukraine R. Szymczak and M. Baran

Institute of Physics, Polish Academy of Sciences, Al. Lotnikow 32/46, PL-02-668 Warsaw, Poland H. Schmid

University of Geneva, Department of Inorganic, Analytical and Applied Chemistry, CJ-1211 Geneva 4, Switzerland ͑Submitted August 3, 2001͒ Fiz. Nizk. Temp. 27, 1208–1213 ͑September–October 2001͒

A study of the magnetization of the antiferromagnetic magnetoelectric crystal LiCoPO4 as a function of temperature and the strength of a magnetic field oriented along the antiferromagnetic vector reveals features due to the presence of a weak ferromagnetic moment. The value of the magnetic moment along the b axis at 15 K is approximately 0.12 G. The existence of a ferromagnetic moment can account for the anomalous behavior of the magnetoelectric effect observed previously in this crystal. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1414584͔

Cobalt lithium orthophosphate is a well-known magne- have been studied, for which it is necessary to apply mag- toelectric crystal of the family of orthorhombic antiferromag- netic and electric fields simultaneously in order to prepare a nets with the olivine structure and having the general for- single-domain state. The behavior of the magnetoelectric ef- ͑ ϭ 2ϩ 2ϩ 2ϩ 2ϩ mula LiMPO4 where M Fe ,Mn ,Co ,Ni ). It has fect observed by the authors of Refs. 1–3 could be due to the been attracting attention because it has large values of the existence of a weak ferromagnetic moment in the crystal. ␣ ␣ 1,2 constants xy and yx of the linear magnetoelectric effect However, previous studies of the magnetic properties of both 7,8 9 and unusual and as yet unexplained behavior of the magne- polycrystalline and single-crystal LiCoPO4 have not de- toelectric effect in a magnetic field.2–4 The magnetoelectric tected weak ferromagnetism. effect in this crystal is the largest among the compounds of The creation of a homogeneous magnetic state in a com- 3d elements. The crystal structure of LiCoPO4, like that of pensated antiferromagnet in the presence of only a magnetic the other lithium phosphates of transition elements of the field may also be caused by quadratic ͑in the field͒ magneti- olivine family, has a symmetry described by the space group zation effect.10 In this case the magnetic field induces in the 16 ͑ ͒ Pnma (D2h) Refs. 5,6 . In this structure the unit cell (a crystal a magnetic moment that is even with respect to the ϭ10.20 Å, bϭ5.92 Å, cϭ4.70 Å͒ contains four formula field strength. In antiferromagnetic ͑AFM͒ states with oppo- units, and the magnetic ions are crystallographically equiva- sitely directed sublattice moments, oppositely directed mag- lent and occupy four c positions. According to the results of netic moments will be induced. Therefore, when a magnetic neutron-diffraction studies7 carried out on polycrystalline field is applied in a certain direction, the energy of the col- samples of LiCoPO4, upon antiferromagnetic ordering (TN linear antiferromagnetic domains will be different. When the ϭ21.9 K͒2 the number of formula units in the unit cell re- difference of the energies of the antiferromagnetic domains, mains unchanged (zϭ4), and the magnetic moments of the which varies in proportion to H3, reaches a threshold value Co2ϩ ions are collinear and directed along the b axis, com- determined by the coercivity of the antiferromagnetic do- pletely compensating each other. The magnetic structure main wall, the crystal will go to a single-domain state or a of the crystal is described in a collinear four-sublattice magnetization reversal of its antiferromagnetic state will 445 Ј occur.11 A quadratic magnetization effect is allowed by sym- model with the Shubnikov symmetry group Sh62 (Pnma ) ͑Ref. 7͒. metry only in AFM crystals which are not symmetric with ͑ In studying the magnetoelectric effect in LiCoPO4,it respect to the operation of anti-inversion or complete inver- 2,3 ͑ ͒ ¯Јϭ¯ Ј Ј was found that for preparation of a homogeneous single- sion :1 1•1 . However, the group mmm that has been ͒ domain antiferromagnetic state of the crystal, as must be established for LiCoPO4, although it does not contain the done in order to measure the magnetoelectric effect, it is operation of spatial inversion, does have a center of anti- sufficient to decrease the temperature of the sample from inversion. Consequently, the quadratic magnetization and Ͼ Ͻ ͑ ͒ T TN to T TN in a magnetic field H oriented along the B weak ferromagnetic moment WFM should be forbidden in axis, or to apply a sufficiently strong magnetic field along the LiCoPO4. Ͻ 12 b axis at temperatures T TN . This is atypical for all of the Magnetooptic studies of LiCoPO4 have revealed new compensated antiferromagnetic magnetoelectric crystals that features of the behavior of this antiferromagnetic crystal in a

1063-777X/2001/27(9–10)/4/$20.00895 © 2001 American Institute of Physics 896 Low Temp. Phys. 27 (99–10), September–October 2001 Kharchenko et al.

FIG. 1. Temperature dependence of the magnetization of the LiCoPO4 crys- ʈ ͑ ͒ tal in an external magnetic field H b with a field strength of 1 T a and 0.05 FIG. 2. Magnetization of the LiCoPO4 crystal as a function of the internal ͑ ͒ ϭ ͑ ͒ ʈ ͑ ͒ T b ; the M(T) curves near the Ne´el temperature at H 0.05 T c .FC magnetic field Hint in an external magnetic field H b a ;theM(H)/H (Hϭ1T͒ — the value of the magnetic field in which the sample was curves at crystal temperatures of 15 K ͑b͒ and 21.3 K ͑c͒. cooled.

a͒ In the temperature interval 5–10.5 K the projection of magnetic fields. It was found that the birefringence of lin- the magnetic moment of the sample on the magnetic field early polarized light induced by a magnetic field Hʈb is com- direction is negative. With increasing temperature its abso- parable to the spontaneous magnetic linear birefringence at a lute value initially increases and then, at TϾ8 K, decreases value of the external field much smaller than the value of the to zero, changes sign, and increases monotonically with tem- ͑ ͒ effective exchange field. This property suggests the presence perature out to the neighborhood of Tt see Fig. 1b and 1c . ͒ ϭ of transverse projections of the magnetic moments in the b At a temperature Tt 20.9 K there is a jump in the crystal and, hence, a noncollinear magnetic structure in it. magnetization ͑occurring in a single interval between experi- Given this situation, it is advisable to make highly sen- mental points, which in this part of the curve is 0.1 K͒. ͒ ͑ ͒ sitive measurements of the magnetization of the LiCoPO4 c Near the Ne´el temperature at 21.6 K see Fig. 1c a crystal. In this paper we report the results of measurements weak but clearly registered peak is observed. The increase of of the magnetization M as a function of temperature and the magnetization with further increase in temperature begins Ͼ magnetic field strength. All of the measurements were made only at T TN . with a SQUID magnetometer ͑Quantum Design MPMS-5͒. All of these anomalies can be explained by assuming The sample studied had a mass of 7.46 mg and was in the that the sample in the initial state had a ferromagnetic mo- form a parallelepiped with dimensions of 0.96ϫ1.22ϫ1.76 ment directed oppositely to the direction of the applied mag- mm. netic field. Thus the sample was initially cooled in ‘‘zero’’ The experimental results are presented in Figs. 1–3. Fig- field ͑ ZFC͒ and then a field Hϭ0.05 T, in which the mea- ure 1 shows the curves of the temperature dependence of the surements were made, was applied. The residual field of the magnetization obtained in magnetic fields H of 1.0 and 0.05 superconducting solenoid can be as high as 0.002 T. The T. The behavior of the magnetization in the 1.0 T field is direction of that field could accidentally be directed opposite similar to that which was observed previously in Ref. 9 at to the direction of the applied magnetic field in the M(T) Hϭ1.2 T. The temperature dependence of M is similar to the measurement. It follows from the trend of the M(T) curve typical behavior for 2D antiferromagnets in which the inter- that the orientation of the spontaneous moment, directed op- action between magnetic ions in the plane is dominant over posite to the field, persists to a temperature Tt , at which the the interaction between ions belonging to neighboring ‘‘switching’’ of the antiferromagnetic state of the sample oc- planes.13 For M(T) there is a broad hump, the top of which curred. In a narrow temperature interval 21–21.9 K the spon- ͑ ϳ ͒ lies above TN at T 25 K . taneous moment is directed along the field. When the tem- In the 0.05 T field the M(T) curves is qualitatively dif- perature approaches TN the spontaneous moment rapidly ferent from that discussed above. It clearly displays the fol- decreases to zero, and a peak appears on the M(T) curve lowing features. ͑see Fig. 1d͒. Low Temp. Phys. 27 (99–10), September–October 2001 Kharchenko et al. 897

ϩ5toϪ5 T and back to ϩ5T.TheM(H) curves are well ϭ ␹ϩ ϩ␤ 3 approximated by the function M(H) M s H H . This function is shown by the solid curve in Fig. 2a for Tϭ15 K. The presence of a nonzero spontaneous moment M s is dem- onstrated by the M(H)/H curves in Figs. 2b, 2c. The posi- tion of the experimental points obtained at Tϭ15 K are well ␹ϩ␤ 2ϩ ϭ described by the function H M s /H, where M s Ϯ0.12 G, ␹ϭ5.06 G/T, and ␤ϭ3.2ϫ10Ϫ2 G/T3. They are shown by the solid and dotted curves in Fig. 2. Because of the nonzero value of M s ,asH goes to zero ϭ␹ϩ␤ 2ϩ the quantity M(H)/H H M s /H increases by a hy- perbolic law, M s /H changes sign when the direction of the field is switched, and then it decreases by the same law. One can see that when the field reaches certain threshold values ϩϭϩ ϪϭϪ ͒ (Hc 1.7 T and Hc 3.8 T a jumplike transition of the experimental points from one branch of the M(H)/H curve to the other occurs. The M(H)/H curve obtained at 21.3 K differs from that described above in that the sign of M(H)/H in the field interval from 5 to 0.1 T is always positive. Only in the field interval from Ϫ0.1 to 0.1 T is the behavior of M(H)/H at 21.3 K sign-varying and similar to that observed at 15 K in a wider field interval. This means that at that temperature the spontaneous moment M s can be directed counter to the field only in a narrow field interval. The hysteresis of the magnetization of the sample is more clearly demonstrated in Fig. 3. In Fig. 3a the part of the curve shown in Fig. 2a is shown in an enlarged scale in the field interval from Ϫ0.1 to ϩ0.1 T. It is seen that a linear extrapolation of M(H)toHϭ0 near zero gives a nonzero spontaneous magnetization for both temperatures. The field dependence of that part of the magnetic moment of the sample which is not due to a simple field-independent sus- ceptibility, i.e., M(H)Ϫ␹(H), is shown in Fig. 3b. The hys- teresis loop is seen against the background of a cubic depen- dence. It is even more clearly seen after subtraction of the cubic contribution ␤H3 ͑Fig. 3c͒. The revealed hysteresis of the magnetization convincingly attests to the presence of a weak spontaneous magnetic moment in the sample. Its value, ϭ ϩϪ Ϫ determined as M s (M s M s )/2, is 0.12 G at 15 K and around 0.015 G at 21.3 K. Other features can also be seen in Fig. 3c. The most intriguing of them is the change in magnetization in the field interval from Ϫ0.1Ttoϩ0.1 T, which has a diamagnetic character. Such behavior of the sample is unusual. It brings to mind ‘‘superdiamagnetism,’’14,15 which is allowed by FIG. 3. Hysteresis of the magnetization in the LiCoPO4 crystal: the M(H) curve in the neighborhood of zero field at temperatures of 15 and 21.3 K ͑a͒; symmetry in this antiferromagnetic magnetoelectric crystal, the M(H) curve after subtraction of the linear contribution ␹H; the dashed where the symmetry of the leading magnetic ordering admits curve shows a dependence of the type ␤H3 ͑b͒;theM(H) curve after the existence of a toroidal moment.4 In the presence of a subtraction of the linear and cubic contributions: M(H)Ϫ(␹Hϩ␤H3) ͑c͒. density gradient of the toroidal moment ͑e.g., near defects of the crystal͒ it is possible to have a diamagnetic response of the system to a magnetic field.15 However, before drawing Further confirmation of the presence of a weak ferro- any conclusions about the causes of the observed feature it is magnet moment in the LiCoPO4 crystal was obtained in a necessary to carry out special experiments and to eliminate study of the field dependence of the magnetization at differ- possible artifacts. ent temperatures. Figure 2 shows the M(T) curves obtained As to the causes of the weak ferromagnetism ͑or, more at 15 and 21.3 K. The temperature 21.3 K was chosen so as precisely, weak ferrimagnetism͒, there are several possible to lie in the narrow temperature interval between the jump mechanisms for its appearance in the LiCoPO4 crystal. Using and the peak on the M(T) curve for Hϭ0.05 T ͑see Fig. 1c͒. the well-known methods of constructing invariants of the Figure 2a shows the field curves of the magnetization, thermodynamic potential,16 one can see that although the obtained for a completely cycle of variation of the field from usual second-order weak-ferromagnetic invariants of the type 898 Low Temp. Phys. 27 (99–10), September–October 2001 Kharchenko et al.

M iL2y are forbidden in LiCoPO4, the Pnma symmetry of an approximation of the M(H) curves by third-order poly- the crystal allows fourth-order invariants containing the nomials does give corrections quadratic in H which have product M yL2y . different signs for the two antiferromagnetic states, they are These invariants are the following: M yL2yL1yL3y , too small. Their small value and substantial difference leaves M yL2yL1xL3x , M yL2yL1zL3z , M yL2yL1xL2z , uncertainty as to whether they correspond to a physical effect M yL2yL1zL2x , M yM zL2yL2z , M yM zL2yL3x , or are the result of a small systematic error and scattering in M yM xL2yL2x , and M yM zL2yL3z . In addition, ‘‘gradient’’ the experimental data. invariants are allowed, which can lead to the formation of a The authors thank V. I. Fomin and V. I. Kut’ko for a modulated magnetic structure. Among them we mention the helpful discussion of the results. Ϫ 4 second-order invariant (M ydL2y /dx L2ydMy /dx). Since *E-mail: [email protected] the projection of the antiferromagnetic vector L2y transforms in the same way as the projection of the toroidal moment Tz ͑Ref. 4͒, the corresponding homogeneous and inhomoge- 1 neous invariants with a toroidal moment are also possible. At M. Marcier, J. Gareyte, and E. F. Bertaut, C. R. Acad. Sci. Paris B 264, 979 ͑1967͒. the present time it is not possible to answer the question of 2 J.-P. Rivera, Ferroelectrics 161, 147 ͑1994͒. which mechanism gives rise to the weak ferromagnetism in 3 H. Wiegelmann, Magnetoelectric Effects in Strong Magnetic Fields,PhD Thesis, University of Konstanz, Konstanzer Dissertationen, Vol. 461, LiCoPO4. We can only say that the nonmonotonic behavior Hartung-Gorre, Konstanz ͑1995͒. of the spontaneous magnetic moment on heating of the 4 I. Kornev, M. Bichurin, J.-P. Rivera, S. Gentil, H. Schmid, A. G. Jansen, sample, which is attested to by the behavior of the magneti- and P. Wyder, Phys. Rev. B 62, 12247 ͑2000͒. zation of the crystal in a field of 0.05 T, indicates the pres- 5 R. E. Newham and M. J. Redman, J. Am. Ceram. Soc. 48,547͑1965͒. 6 ence of competing mechanisms. A similar dependence of the F. Kubel, Z. Kristallogr. 209,755͑1994͒. 7 R. P. Santoro, D. J. Segal, and R. E. Newnham, J. Phys. Chem. Solids 27, weak-ferromagnetic moment has been observed in crystals of 1192 ͑1966͒. antiferromagnetic vanadates, in particular, yttrium 8 A. Goni, L. Lezama, G. E. Barberis, J. L. Pizzarro, M. I. Arriortua, and vanadate.17 It must also be noted that the features of the T. Rojo, J. Magn. Magn. Mater. 164, 251 ͑1996͒. 9 J.-P. Rivera and J. Korean, Phys. Soc. 32, S1855 ͑1998͒. behavior of the magnetic linear birefringence in a cyclically 10 ͑ ͒ 12 G. Gorodetsky and D. Treves, Phys. Rev. A 135,97 1964 ;G.Goro- varying magnetic field implies that LiCoPO4 must have an detsky, B. Sharon, and S. Shtrikman, Solid State Commun. 5,739͑1967͒. incommensurate modulated structure, if not spontaneous 11 A. V. Bibik, N. F. Kharchenko, and S. V. Petrov, Fiz. Nizk. Temp. 15,1280 then magnetic-field induced. Depending on the mechanism ͑1989͓͒Sov. Phys. JETP 15, 707 ͑1989͔͒; N. F. Kharchenko, R. Szymczak, ͑ ͒ giving rise to the weak ferromagnetism, the magnetic point and M. Baran, J. Magn. Magn. Mater. 140, 161 1995 . 12 M. F. Kharchenko, O. V. Miloslavska, Yu. M. Kharchenko, H. Schmid, group of the crystal may be one of the following noncen- and J.-P. Rivera, Ukr. J. Phys. Opt. 1,16͑2000͒. troantisymmetric groups: m(Ќy) ͑this group admits a ‘‘lon- 13 M. F. Sykes and M. E. Fisher, Physica 28,919͑1962͒. gitudinal’’ weak-ferromagnetic structure M L L L ); 14 V. L. Ginzburg, A. A. Gorbatsevich, Yu. V. Kopayev, and B. A. Volkov, y 2y 1y 3y ͑ ͒ 2Ј(ʈx) ͑which admits the structures M L L L , Solid State Commun. 50, 339 1984 . y 2y 1x 2z 15 A. A. Gorbatsevich and Yu. V. Kopayev, Ferroelectrics 161,321͑1994͒. ͒ Ќ 16 M yL2yL1xL3x , M yM zL2yL2z , and M yM zL2yL3x , mЈ( z) E. A. Turov, Physical Properties of Magnetically Ordered Crystals ͓in ͑ ͔ ͑ ͒ which admits the structures M yL2yL1zL3z , M yL2yL1zL2x , Russian , Izd. Akad. Nauk SSSR, Moscow 1963 . M M L L , and M M L L ). 17 Y. Ren, T. T. M. Palstra, D. I. Khomskii, and G. A. Sawatsky, Phys. Rev. y x 2y 2x y z 2y 3z ͑ ͒ The presence of a quadratic magnetization effect in the B 62, 6577 2000 . ʈ LiCoPO4 crystal for H b remains open question. Although Translated by Steve Torstveit LOW TEMPERATURE PHYSICS VOLUME 27, NUMBER 9–10 SEPTEMBER–OCTOBER 2001

Orthogonal-to-unitary ensemble crossover in the electronic specific heat of metal nanoclusters S. I. Mukhin* and O. V. Kurapova

Theoretical Physics Department, Moscow Institute for Steel and Alloys, Moscow, Russia I. I. Moiseev and M. N. Vargaftik

N. S. Kurnakov Institute of General and Inorganic Chemistry, Russian Academy of Sciences, Moscow, Russia Y. E. Volokitin, F. L. Mettes, and L. J. de Jongh

Kamerlingh Onnes Laboratory, Leiden University, Leiden, The Netherlands** ͑Submitted August 27, 2001͒ Fiz. Nizk. Temp. 27, 1214–1220 ͑September–October 2001͒ We present a theoretical and experimental study on the influence of a magnetic, field on the energy- level statistics in metal nanoparticles. Based on the random-matrix theory, a gradual field- induced crossover behavior is predicted from the orthogonal to the unitary ensemble. Experimental data of the electronic specific heat of metal nanoparticles for different field sin the quantum size temperature regime compare favorably with these theoretical ͑analytical͒ predictions. © 2001 American Institute of Physics. ͓DOI: 10.1063/1.1415943͔

1. INTRODUCTION The term ‘‘geometrical’’ used above is based on the fact that the Jacobian relates volumes in the two equivalent abstract A large number of papers have been devoted to the ex- spaces of the random variables, thus reflecting their geo- perimental and theoretical investigation of the thermody- metrical structure: namic properties of small metal particles ͑nanoclusters͒.In order to test the predictions based on the random-matrix ͑ ͒d␮͑ ͒ϭd␮͑U͒ͩ P͕͑E ͖͒J͕͑E ͖͒͟ dE ͪ . ͑4͒ ͑ P H H n n i theory of the energy-level statistics for reviews see Refs. i 1–4͒. The discrete character of the energy spectrum of a nanocluster can be regarded as a direct manifestation of the Here the volume element d␮(H) is the ‘‘natural’’ volume quantum size effect. The Wigner–Dyson5,6 random ensemble measured in the space of the Hermitian matrix elements ␮ ͟ description relies upon the functional P (H): Hij , which factorizes into the product d (U) idEi of the ␮ ͟ volume elements d (U) and idEi in the subspaces of the P ͑H͒ϰexp͓Ϫv Tr∨͑H͔͒. ͑1͒ eigenvectors and eigenvalues of the matrix H. As is appar- ent from Eq. ͑3͒ and ͑4͒, the character of the statistical dis- 7 This functional is chosen to describe the probability distribu- tribution of the level-spacings depends only on the index v, tion for the random ensemble of NϫN Hamiltonian matrices which counts the number of degrees of freedom in the matrix of the electrons in the nanoclusters. An ensemble is called elements Hij . This number reflects the symmetry of the Gaussian if ∨(H)ϰH.2 This distribution is successfully system in question and can only take the values 1, 2 or 4 for used in the absence of strong electron-electron the real, complex or real quaternion matrix elements, respec- correlations.1,2 In the Gaussian distribution the argument of tively. The matrix elements may be chosen real when time- the exponential becomes a simple sum over all the matrix reversal symmetry exists, while they become imaginary 2ϭ͚ ͉ 2 ͉ϭ͚ 2 when this symmetry is broken, e.g., by an external magnetic elements: Tr H ij H iE , so that it does not con- ij i field or magnetic impurities. In time-reversal symmetric sys- tribute to the correlations between the levels ͕Ei͖, because tems with broken spin-rotation symmetry, e.g., as a conse- the probabilities for the different E j’s factorize: quence of spin-orbit interaction, the matrix elements are real quaternions. The details of the interactions inside the system ͑ ͒ϰ ͓Ϫ 2͔ϰ ͓Ϫ 2͔ϭ ͕͑ ͖͒ P H exp v Tr H ͟ exp vEi P En . do not change these general symmetries, and hence, the spec- i tral correlations posses universality. The corresponding ͑ ͒ 2 Gaussian ensembles of random matrices, depending on their symmetry, are called orthogonal (vϭ1), unitary (vϭ2) or Therefore, the spectral correlations in the Gaussian ensemble ϭ 1 symplectic (v 4) ensembles respectively, or in the abbrevi- are of a purely geometrical nature, since they follow solely ated form: GOE, GUE or GSE. Thus, the symmetry of the ͕ ͖ 3 from the Jacobian J( En ): Hamiltonian matrices of the clusters changes either from GOE to GUE, or from GSE to GUE upon application of an ͒ϭ ͉ Ϫ ͉v ͑ ͒ external magnetic field. We remark that the consequences of J͕͑En͖ ͟ Ei E j . 3 iϽ j random perturbations on the energy level spectra of an as-

1063-777X/2001/27(9–10)/6/$20.00899 © 2001 American Institute of Physics 900 Low Temp. Phys. 27 (99–10), September–October 2001 Mukhin et al. sembly of small metal particles and the ensuing thermody- Here the variance v2 determines the mean level spacing ␦ namic properties were first investigated by Kubo,8 using the ϭ␲v/ͱN in the vicinity of the Fermi energy of the cluster in Poisson distribution for the energy level statistics. Shortly the limit Nӷ1 and ␣Ӷ1. As was explained above, the dou- afterwards, Gor’kov and Eliashberg9 pointed out the rel- bling of the dimensionality of the phase-space of the off- evance of the Wigner–Dyson formalism for small metal par- diagonal matrix elements leads16 in turn to the doubling of ticles and demonstrated the profound influence of the level the exponent v in the eigenvalue ͑level͒ distribution function ϭ ϭ repulsion on the low-temperature thermodynamics of the P(͕En͖), from v 1 in GOE to v 2 in GUE: small metal particles in the two limiting cases of GOE and GUE. A more detailed consideration of the metal-cluster ͒ ͒ϰ ͉ Ϫ ͉ 10 PGOE͕͑En͖ JGOE͕͑En͖ ͟ Ei E j ͟ thermodynamics in these two limits was later made in. iϽ j k Theoretical justification of the Wigner–Dyson statistics was 11 ͑ ͒ provided in for the case of diffusive nonballistic chaotic ϫ ͓ 2͔→ ͉ Ϫ ͉2 exp Ek ͟ Ei E j ͟ electron motion inside the cluster. The most recent theoreti- iϽ j k cal achievements, which include justification of the Wigner– ϫ ͓Ϫ 2͔ϰ ͕͑ ͖͒ ͕͑ ͖͒ ͑ ͒ Dyson statistics for the nonintegrable chaotic systems, ‘‘bal- exp 2Ek PGUE En JGUE En . 8 listic billiards,’’ with random scattering at the boundaries, 1 The subtle feature of this crossover is that the doubling of v and summarized in. 2 In this note we present a preliminary report on detailed is actually energy-dependent. The GOE-to-GUE transition ͉␦ ͉у ␦ theoretical predictions,12 together with a comparison to re- is completed on the energy scale E if Ei E, where Ei is ␣ cently obtained experimental data,13–15 for the specific heat the energy shift due to the ‘‘perturbation’’ i A in the ͑ ͒ behavior of metal nanoclusters in the crossover regime in- Hamiltonian 5 : duced by an external magnetic field. The crossover manifests A2 itself as a gradual transition from the orthogonal to unitary ␦ ϭ␣2 ij ͑ ͒ Ei ͚ Ϫ . 9 Gaussian distribution of the electronic energy levels ͑GOE- iÞ j Ei E j to-GUE transition͒ of the nanoclusters. We remark that these measurements represent the first experimental observation of Simultaneously, the ‘‘high-energy’’ part of the spectrum on ӷ␦ the presence of the quantum-size effect in the electronic spe- the scale E Ei remains distributed according to the upper cific heat and susceptibility of metal nanoparticles. line of Eq. ͑8͒ with vϭ1, as it would be in the case of an Mathematically, the basic change resulting from the orthogonal ensemble. Therefore, measuring the thermody- GOE-to-GUE crossover lies in the appearance of the ͑ran- namic properties of the nanoparticles in the external mag- dom͒ imaginary part of the electron Hamiltonian matrix due netic field may give an experimental verification of the rela- to the breaking of the time-reversal symmetry by the external tion ͑8͒ and of the gradual nature of the GOE-to-GUE magnetic field. This phenomenon was modeled by Pandey transition, provided that these properties depend on the dif- and Mehta16 by the addition of a random antisymmetric real ferent energy scales at the different temperatures. matrix A with imaginary weight i␣ to the real symmetric On basis of the above discussion we expect the specific Hamiltonian matrix H0 : heat of the nanoclusters to be a relevant thermodynamic characteristic for the observation of the GOE-to-GUE cross- ϭ ϩ ␣ ͑ ͒ H H0 i A. 5 over in magnetic field. In order to investigate this problem quantitatively, we consider an ensemble of metal particles ␣ ⌽ The parameter is proportional to the magnetic flux with half of the clusters having an even number of electrons through the system, i.e., through the nanoparticle, so that the and the other half having an odd number. Our calculation ␣ ⌽ relation between and depends on the geometry of the was done12 in the low temperature limit, TӶ␦, so that only a ͑ ͒ particle and on the ratio of its size radius R to the electron few electronic levels need to be explicitly considered. Taking mean fee path I. For a ballistic sphere (RӶI) with diffuse 2 into account all the different possibilities of the formation of boundary scattering of electrons this relation is the lowest excited energy states, one finds the following ex- pressions for the low-temperature partition functions of the e⌽ 2 បv 8␲ ␣2ϭͩ ͪ F ͑ ͒ even and odd clusters, Z and Z respectively: N , 6 even odd h R␦ 45 Z ϭ͑1ϩ2eϪ␤␦ cosh͑h␤͒͒͑1ϩ2eϪ␤␦͒ϩO͑eϪ2␤␦͒, where is the Fermi velocity, ␦ is the mean level spacing at even vF ͑10͒ the Fermi energy of the nanoparticle, and N is the matrix dimension indicating the total number of the single-electron Z ϭ2 cosh͑h␤/2͒͑1ϩ2eϪ␤␦͒ϩO͑eϪ2␤␦͒, ͑11͒ states taken into account in the model of the nanoparticle. odd The addition of an independently randomly distributed ma- where ␤ϭ1/T is the inverse temperature in energy units, i.e., trix A to the ͑random͒ real symmetric matrix H , effectively ϭ ϭ ␮ ␮ 0 taking kB 1. Also h g BH, where B is the Bohr magne- doubles the number of degrees of freedom in the elements of tron, gϭ2 is the Lande´ factor, and H is the external magnetic the Hamiltonian matrix H, so that the Gaussian distribution field. of H takes the form Using Eqs. ͑10͒ and ͑11͒ we have for the free energy F

2 2 H0ij Aij 1 P ͑H͒ϰexpͩ Ϫ͚ ϩ ͪ . ͑7͒ FϭϪ T͑ln Z ϩln Z ͒, ͑12͒ i, j 4v2 4v2 2 even odd Low Temp. Phys. 27 (99–10), September–October 2001 Mukhin et al. 901 where the coefficient 1/2 represents the fifty/fifty probability non-vanishing probability of even the smallest level spac- to find an even/odd cluster in the macroscopically large as- ings. When these spacings are much less than the average sembly of otherwise identical metal particles. distance between the levels ͑␦͒, they will make essential con- The specific heat CV , magnetization M, and magnetic tributions to the specific heat at the lowest temperatures, susceptibility ␹ can be derived from the well-known rela- TӶ␦, so that the energy gap due to the quantum size effect tions becomes a pseudo-gap. Then, the explicit averaging over the appropriate level distribution is essential to describe the data. 2Fץ Fץ 2Fץ ϭϪ ϭϪ ␹ϭϪ ͑ ͒ It proved to be possible to perform such an averaging CV T , M , . 13 ,H2 analytically12 for the case of the low enough temperaturesץ Hץ T2ץ where only the lowest excited states of the clusters should be In the simplest equidistant model for a metal nanoparticle, important. The approximation used here consists in the aver- due to Fro¨hlich,17 in which all the levels are at a constant aging of the low-temperature thermodynamic functions interlevel spacing ␦, these thermodynamic functions are A(␦,T,H), found for the equidistant model, with the two- given by the formulas: level correlation function R2 . For this purpose we substitute 1 1 eϪ␦␤ an equidistant level spacing ␦ by a random variable ␧ and ϭ ͑ ␤2͒ ϩ ͑␦␤͒2 CV h 2 then perform an integration over this variable: 8 cosh2͑h␤/2͒ ͑1ϩ2eϪ␦␤͒2 ϱ ␧ ␧ 2 Ϫ2␦␤ϩ͑ Ϫ␦͒2 ␤͑hϪ␦͒ϩ͑ ϩ␦͒2 Ϫ␤͑hϩ␦͒ ¯ ͑ ␦͒ϭ ͵ ͩ ͪ ͑␧ ͒ ͩ ͪ ͑ ͒ 1 4h e h e h e A T,H, R2 ␦ A ,T,H d ␦ . 17 ϩ ␤2 , 0 2 ͑1ϩ2eϪ␦␤ cosh͑h␤͒͒2 ͑14͒ Let the level distribution function pn(x) define the prob- ability to find a spacing x between two energy levels, while 1 sinh͑h␤͒ there are n other energy levels being located between those ϭ ␮ ͫ ͑ ␤ ͒ϩ ͬ ͑ ͒ M g B tanh h /2 , 15 ␧ 4 e␦␤ϩ2 cosh͑h␤͒ two. Here we normalize the random level spacing by the average interlevel spacing ␦ near the Fermi-level of a cluster: 1 2ϩe␦␤ cosh͑h␤͒ xϭ␧/␦. Then, the two-level correlation function is defined ␹ϭ␤͑ ␮ ͒2ͫ ϩ ͬ g B . as follows: 8 cosh2͑h␤/2͒ ͑e␦␤ϩ2 cosh͑h␤͒͒2 ͑16͒ ϱ ͒ϭ ͒ ͑ ͒ R2͑x ͚ pn͑x . 18 It is obvious from the above expressions that in the equidis- 0 tant model there is effectively a gap in the density of states at Ӷ → the Fermi-level, leading to an exponentially vanishing spe- At small enough x 1, we have R2(x) p0(x), and in ӷ → cific heat at low enough temperatures, i.e., TӶ␦. The oppo- the limit x 1, R2(x) 1. From the work of Pandey and 16 site extreme to the current model is the Poisson level distri- Mehta, using also relation Eq. ͑6͒͑see Ref. 2͒, we have an Ϫ ϰ Ϫ͉ Ϫ ͉ ␦ analytical expression for R for an arbitrary external mag- bution P0(Ei E j) exp͓ Ei Ej / ͔, corresponding to no 2 gap at all, which was used by Kubo.8 Unlike in the distribu- netic field H: tions ͑8͒, the Poisson’s distribution entails a finite probability sin͑␲x͒2 1 ␲ for the two levels to ‘‘stick together,’’ thus neglecting the R ͑x͒ϭ1Ϫͩ ͪ ϩ ͵ k sin͑kx͒ 2 ␲ 2 level repulsion. Hence, the pseudo-gap in the density of x ␲ 0 states at the Fermi-energy, caused by the off-diagonal ran- ϱ 2 2 sin͑zx͒ 2 2 dom matrix elements between different electronic states in ϫe2k ␳ dk͵ eϪ2z ␳ dz, ͑19͒ the cluster, disappears in the Poisson model. ␲ z ␳ϭ ␮ ␦ where 1.15 BH/ for a spherical particle and where bal- 2. AVERAGING OVER THE LEVEL DISTRIBUTION IN THE listic electron motion inside the particle is supposed.12 Then, CROSSOVER REGION ␳ϭ Ӷ ϳ without magnetic field, 0, R2(x 1) x, while in the ␳ӷ ␲ Ӷ ϳ 2 Our aim here is to obtain more realistic expressions for strong field limit, 1/ , R2(x 1) x . The probability to the thermodynamic functions of the metal nanoclusters than find two levels close together, xӶ1, decreases with the field those given in Eqs. ͑14͒, ͑15͒, and ͑16͒, which were derived from ϳx to ϳx2. This effect is called the ‘‘interlevel repul- for an equidistant model of the energy levels. Actually, for sion’’ in a magnetic field and causes a change of the low- small particles one usually assumes that minor perturbations temperature dependence of the field-independent contribu- such as surface irregularities ͑even of an atomic scale͒ will tion to the electronic specific heat ͑second term on the right ͑ ͒ ͑ ͒͒ GOEϰ 2 be sufficient to make the level distribution random. Here, we hand side rhs of Eq. 14 from Cel T for GOE (H ϭ GUEϰ 3 ␮ ӷ␦ 9,10 will suppose the random distribution to obey the Wigner– 0) to Cel T for GUE ( BH ). Incidentally, in Dyson Gaussian ensemble statistics. Indeed, the published case of metal nanoparticles with strong spin-orbit coupling experimental data14 for the temperature dependences of the ͑GSE level-statistics͒, this contribution to specific heat is ex- GSEϰ 5 10 specific heat of the metal nanoparticles clearly showed the pected to behave as Cel T at low enough temperatures. ͓ inapplicability of the equal level spacing model, in which On the other hand, the spin-flip contribution to Cel incorpo- case the specific heat becomes exponentially small at the rated in the third term on the rhs of Eq. ͑14͔͒, after averaging Ӷ␦ temperatures T . Comparison of the experimental curves with the function R2(x), brings T-linear contribution to the 14 in with the theory leads to the conclusion that the experi- electronic specific heat (Cel), which masks the crossover in mental behavior has to be described by a model based on a the temperature dependence described above. This later fact, 902 Low Temp. Phys. 27 (99–10), September–October 2001 Mukhin et al.

FIG. 2. Electronic specific heat ͑in zero field͒ of the metal nanoclusters discussed in the text. Data were scaled on the theoretical prediction ͑solid curve͒ with the average energy level spacing ␦ as the adjustable parameters. The transition from bulk-like behavior (ϰT) to quantum-size behavior (ϰT2) can be clearly seen. FIG. 1. Calculated temperature dependences of the electronic specific heat ͑ ͒ Cel and its different constituents lower panel , and of their logarithmic derivatives ͑upper panel, for h/␦ϭ0.1 only͒ for a random Gaussian en- semble of metal clusters at two different values of the external magnetic interval, while the crossover from ␣ϭ3 via ␣ϭ2to␣ϭ1in ͑ ␮ ␦ ␦ ͑ field H in units g BH/ ). The temperature is expressed in units T/ put- the higher temperature region is quite pronounced in the up- ϭ ting kB 1). per panel of Fig. 1. In the higher magnetic field hϭ1␦ the Schottky peak had moved to higher temperatures where it is superimposed on the much larger even cluster contribution to as far as we know, was not considered in the previous works, 3 9,10 the specific heat and so it does not mask substantially the T see. Below we present the calculated dependences of CV term in the total Cel as in the weakest field case. Neverthe- on the temperature and the external magnetic field H and ϰ ␦ less, the spin-flip contribution, which is TR2(h/ ), still compare them with the recently obtained experimental re- 3 sults. masks the T GUE term in the low temperature interval: T р͕h,␦͖. In the even stronger magnetic field hϭ1.62␦ ͑not shown in Fig. 1͒ practically the whole low temperature inter- 3. DISCUSSION val TӶ␦ is occupied by the unitary ensemble contribution to GUE ϰ 3 sfϰ As a main result to be concluded from the calculated the specific heat Cel (T) T . The spin-flip term Cel T is ␦ ␦ dependences shown in Fig. 1, we point out that the GOE-to- again the strongest one here, as R2(h/ ) saturates at h/ ӷ → GUE transition in magnetic field has a gradual crossover 1: R2 1. Hence, the spin-flip term mimics the ‘‘bulk’’ character. The underlying reason is that the distribution of behavior at low temperatures Tр␦Ͻh. Then, at Tу␦, ␧ Ӷ the level spacings on the ‘‘small energy scale,’’ /h 1, Cel(T), smoothly crosses-over into the ‘‘actual’’ bulk behav- ␦Ӷ ϰ obeys GUE statistics already at small fields h/ 1, while ior, Cel(T) T. The two regions are separated by the the level spacings for large energy, ␧/hӷ1 remain distrib- Schottky contribution, which in this case could be visible ץ ץ uted according to GOE until the field becomes strong only in the ln Cel / ln T vs. In T dependence, but is hardly ␦у ␦Ӷ ͑ ͒ enough, i.e., h/ 1. Thus, in small fields, h/ 1, the tem- noticeable in the Cel(T) curve not shown . Furthermore, perature dependence of the specific heat may be separated based on the above discussion, we would expect a direct ϰ 2 ϰ into three different regions. In the lowest temperature region, crossover from the GOE, Cel(T) T to the bulk Cel(T) T Ӷ Ӷ␦ T h , the electronic specific heat Cel contains the unitary behavior of specific heat in zero magnetic field, where the GUE ϰ ␣ ␣ϭ ensemble part: Cel (T) T , with 3, together with the spin-flip contribution is absent. Indeed, this behavior was sf ϰ ␦ spin-flip contribution: Cel (T) TR2(h/ ). In the intermedi- measured in a series of Pd metal clusters of different sizes ate temperature interval, hӶTӶ␦, the temperature depen- ranging from 2 nm to 15 nm in diameter,13–15 and is shown dence remains the same as in the orthogonal ensemble, i.e., in Fig. 2. In this figure the electronic contributions to the ␣ϭ ϰ 2 with 2: Cel(T) T , because the spin-flip contribution in specific heat are shown, which were obtained at very low this case merely changes the numerical prefactor in front of temperatures (TϽ1K͒, by subtracting the phonon ͑lattice͒ T2. Finally, in the ‘‘high temperature’’ region, Tӷ␦, the contributions from the measured data. The special feature of temperature dependence of the electronic specific heat fol- these metal clusters is that they appear as part of chemically ϰ lows the usual Fermi-liquid bulk behavior: Cel(T) T. Dif- synthesized, molecular compounds, each compound contain- ferent characteristic regions are clearly seen in Fig. 1, where ing metal clusters of a given uniform size embedded in a the contributions to specific heat are plotted separately for stabilizing ligand shell.18–21 The metal cluster compounds the two different magnetic field values as functions of the mentioned here were made by two different groups, namely temperature in units of the average level spacing ␦. As can be that of Profs. Moiseev and Vargaftik in Moscow, and the seen in Fig. 1, in the weakest field, hϭ0.1␦, the Schottky group of Prof. G. Schmid at the University of Essen. The GUE contribution due to the odd clusters masks the Cel (T) compounds indicated by Pd5,Pd7, and Pd8 contain respec- ϰT3 term within a substantial part of the low temperature tively 561, 1415, and 2057 Pd atoms per cluster. There are Low Temp. Phys. 27 (99–10), September–October 2001 Mukhin et al. 903

FIG. 4. Temperature dependence of the electronic specific heat of Pd5 ob- tained after subtraction of the lattice contribution. Dotted curve: equal level spacing model. The other curves follow from the calculations discussed in the text ͑compare Fig. 1͒.

surements, i.e., with the lattice contributions not yet sub- tracted. The latter give a T3 term, which becomes rapidly negligible below 1 K, as can be seen most clearly from the Pd-coll data. The specific heat behavior of the bulk Pd has been included in Fig. 3 as the dotted curves. Comparing this with Pd-coll, one observes that for the latter the linear term is still 20–30% lower than for the bulk, which can be under- stood in terms of a lower average density of states at the FIG. 3. Temperature dependence of the measured specific heat ͑including Fermi-energy in the nanoparticles due to surface effects.22 ͒ phonon contributions for Pd5,Pd7, and Pd-coll. The dotted curves give the Finally, Fig. 4 shows the field-dependence measured for behavior found for bulk Pd. Pd5 on linear temperature scale, with the phonon contribution subtracted. The zero-field data clearly display the T2 depen- dence predicted for the GOE model, in contrast with the two Pd compounds, namely with chemical formula 5 exponential dependence following from the equal level spac- Pd Phen ͑DAc͒ synthesized by Moiseev and Vargaftik 561 60 180 ing model ͑dotted curve͒. The Schottky anomaly found in ͑Pd -M͒, and Pd Phen O synthesized by Schmid 5 561 36 200 low field is ascribed to the odd-electron clusters and is in (Pd5-S) The chemical formulas of Pd7 and Pd8 are ͑ ͒ good agreement with the theoretical prediction ͑also shown Pd Phen bu O and Pd Cinch , respectively. Here, 1415 114 1650 2057 56 in Fig. 1 for h/␦ϭ0.1͒. Cinch stands for cinchonidine, Phen for phenanthroline and ͑ ͒ In conclusion, we have presented theoretical and experi- Phen bu is a phenanthroline derivative. The sample denoted mental results describing the quantum-size effects in the by Pd-coll is a Pd colloid with average particle size of 1.25 thermodynamic properties of nanosize metal clusters with ϫ105 atoms/particle and a small size distribution of 5 – 10%. random distribution of the electronic energy levels in differ- The Pd and Pd metal clusters compounds and the Pd-coll 7 8 ent external magnetic field sin the few-Kelvin temperature were also synthesized by Schmid and coworkers. range. The experimental data show a good qualitative agree- It is worth pointing out that the only adjustable param- ment with theoretical predictions based on the Wigner– eter needed to scale the experimental C vs. T data upon the el Dyson/Mehta–Pandey random-matrix theory. theoretical curve in Fig. 2 is the average energy level spacing We like to thank Prof. G. Schmid for providing the ␦. The values used are given in the figure and indeed vary samples quoted in this study. This work is part of the re- roughly with the inverse volume of the particles, as expected search program of the Stichting voor Fundamenteel Onder- from theory ͑the particle diameters are approximately 2.4, zoek der Materie ͑FOM͒, which is sponsored by the Neder- 3.2, 3.7, and 15 nm for Pd5,Pd7,Pd8, and Pd-coll, respec- ͒ landse Organisatie voor Wetenschappelijk Onderzoek tively . Accordingly, the crossover from quasi-bulk behavior ͑ ͒ ϰ ϰ 2 ␦ NWO . (Cel T) to the quantum-size regime (Cel T ) near T/ ϭ0.2 actually occurs at lower temperature, the larger is the DEDICATION particle. This same trend can also be seen in the experimental field dependence of the specific heat, as shown in Fig. 3. It It is our great pleasure to dedicate this paper to the should be pointed out here that the mere presence of a field memory of Prof. Lev Shubnikov, who was a prominent visi- dependence is in itself a quantum size effect, since the bulk tor of the Kamerlingh Onnes Laboratory in the last century, metal will not show such behavior. Comparing the data of where, together with Prof. W. J. de Haas. He discovered the Pd5,Pd7, and Pd-coll, one observes that the onset tempera- famous effect that bears their names. As witnessed by this ture of the field-dependence is indeed lower, the larger the paper, also recently there have been intensive relations be- particle. We note that the data in Fig. 3 show the raw mea- tween scientists at the Kamerlingh Onnes Laboratory and the 904 Low Temp. Phys. 27 (99–10), September–October 2001 Mukhin et al. former Soviet Union. It is our hope that these fruitful col- 10 R. Denton, B. Muhlschlegel, and D. J. Scalapino, Phys. Rev. B 7,3589 laborations, from which we have had so much profit and ͑1973͒. pleasure, will be an inspiration for future workers to continue 11 K. B. Efetov, Adv. Phys. 32,53͑1983͒. 12 and intensify scientific exchange worldwide. S. I. Mukhin and O. V. Kurapova, manuscript in preparation. 13 Y. E. Volokitin, Ph.D. Thesis Leiden University ͑March, 1997͒. 14 Y. E. Volokitin, J. Sinzig, L. J. de Jongh, G. Schmid, M. N. Vargaftik, and * E-mail: [email protected] and [email protected] I. I. Moiseev, Nature ͑London͒ 384, 621 ͑1996͒. **E-mail: [email protected] 15 F. L. Mettes, Ph.D. Thesis Leiden University ͑in preparation͒. 16 A. Pandey and M. L. Mehta, Commun. Math. Phys. 87,449͑1983͒. 17 H. Frolich, Physica 4, 406 ͑1937͒. 1 T. Guhr, A. Mu¨ller-Groeling, and H. A. Weidenmuller, Phys. Rep. 299, 18 M. N. Vargaftik, V. P. Zagorodnikov, I. P. Stolyarov, I. I. Moiseev, V. A. 189 ͑1998͒. Likholobov, D. I. Kochubey, A. L. Chuvilin, V. I. Zaikovsky, K. Zama- 2 ͑ ͒ C. W. J. Beenakker, Rev. Mod. Phys. 69,731 1997 . raev, and G. I. Timofeeva, J. Chem. Soc. Chem. Commun. 937, 1985 ͑͒. 3 ͑ ͒ W. P. Halperin, Rev. Mod. Phys. 58,533 1986 . 19 G. Schmid, Chem. Rev. 92, 1709 ͑1992͒; G. Schmid, Struct. Bonding 4 ͑ ͒ J. A. A. J. Perenboom, P. Wyder, and F. Meier, Phys. Rep. 78, 175 1981 . ͑ ͒ ͑ ͒ 5 ͑ ͒ Berlin 62, 5285 1985 . E. P. Wigner, Proc. Cambridge Philos. Soc. 47, 790 1951 ; E. P. Wigner, 20 ͑ ͒ in Statistical Theories of Spectral: Fluctuations, edited by C. E. Porter, Clusters and Colloids: From Theory to Applications,G.Schmid Ed. , ͑ ͒ Academic Press, New York ͑1965͒. VCH, Weinheim 1994 . 21 6 F. J. Dyson, J. Math. Phys. 3,1199͑1962͒; F. J. Dyson and M. L. Mehta, Physics and Chemistry of Metal Cluster Compounds,L.J.deJongh͑Ed.͒, ibid. 4,701͑1963͒. Kluwer Academic, Dordrecht ͑1994͒. 7 See C. E. Porter ͑Ed.͒, Statistical Theories of Spectra: Fluctuations, Aca- 22 D. A. van Leeuwen, J. M. van Ruitenbeek, and L. J. de Jongh, Phys. Lett. demic Press, New York ͑1965͒. A 170, 325 ͑1992͒. 8 R. Kubo, J. Phys. Soc. Jpn. 17, 975 ͑1962͒. 9 L. P. Gor’kov and G. M. Eliashberg, Zh. E´ ksp. Teor. Fiz., ͑1965͓͒Sov. This article was published in English in the original Russian journal. Repro- Phys. JETP 21,940͑1965͔͒. duced here with stylistic changes by AIP.