<<

American Mineralogist, Volume 101, pages 1036–1045, 2016

A century of structures: How well do we know them?

Ross J. Angel1,* and Fabrizio Nestola1

1Department of Geosciences, University of Padua, Via Gradenigo 6, I-35131, Padua, Italy

Abstract This decade marks the centenary of the discovery of X‑ray diffrac- tion. The development of as a scientific discipline in which the properties of are understood in terms of their atomic-scale structures has paralleled the development of diffraction crystallography. As diffraction crystallography revealed more precise details of mineral structures, more subtle questions about mineral properties could be ad- dressed and a deeper understanding of the relationship between the two could be attained. We review the developments in X‑ray single-crystal diffraction crystallography over the last century and show how its power to provide fundamental information about the structures of minerals has evolved with the improvements in data quality and the increased technological capac- ity to handle the data. We show that modern laboratory X‑ray diffraction data are of the quality such that mineralogical results are no longer limited by the data quality, but by the physical validity of the refinement models used to interpret the data. Keywords: Crystal structures, X-ray diffraction, site occupancies, anorthite, , , Invited Centennial article

Introduction of the diffraction of X‑rays by crystals proved two things; X‑rays Minerals are, by definition, crystalline. Their structures are can behave as waves, and crystals are periodic arrays of X‑ray composed of atoms arranged in periodically repeating patterns. scatterers. Within a year, the Braggs had used the intensities of the The same atoms in different minerals tend to occupy characteristic X‑ray beams diffracted by crystals to determine the arrangements local environments with specific coordination geometries and of the atoms in crystals of simple minerals such as the zinc blende bond lengths. The arrangements of atoms within minerals and the used in the experiments in Muenchen, and NaCl, and showed that crystal-chemical concepts derived from them successfully explain they were not molecular compounds (e.g., Gruner 1929). The many of the physical, chemical, and thermodynamic properties of subsequent development of mineralogy as a scientific discipline minerals. But how do we know that minerals are built of atoms? in which the properties of minerals are understood in terms of their More than a century ago, there was no proof of the existence of atomic-scale structures (e.g., Gruner 1931; Kieffer and Navrotsky atoms, although it was clear that gases behaved as if composed of 1985) has paralleled the subsequent development of diffraction particles that occupied some finite volume. It had been deduced crystallography. As diffraction crystallography revealed more from the regularity of their external form (but not proven) that precise details of structures, more subtle questions about mineral crystalline materials and especially minerals were probably pe- properties could be addressed and a deeper understanding of the riodic arrays of “some things”; the common deduction was that relationship between the two could be attained. By following crystals were composed of the same “molecular” species that occur the developments over the last century we can understand both in gases, condensed into regular and periodic arrays (Kubbinga the power of diffraction crystallography to provide fundamental 2012). In the first half of 1912 three young men in the Institute information about the structures of minerals, while at the same for Theoretical Physics of the Ludwigs-Maximillians Universi- time appreciating its important intrinsic limitations and how these taet Muenchen carried out an experiment that would confirm the limitations have changed as the technology supporting X‑ray dif- atomistic view of the chemical and mineral world. Inspired by fraction has developed. conversations with P.P. Ewald, Max von Laue had suggested the The development of diffraction experiment to two other research assistants, Friedrich Knipping crystallography and Walter Friedrich. In May 1912 they submitted a manuscript to the Bavarian Academy for Science that reported the diffraction The very first diffraction measurements and structure solutions of X‑rays by crystals (Friedrich et al. 1912). It was known from provided a way for diffraction crystallography to “boot-strap” its optics and physics that diffraction occurs due to constructive and own methodology. Bragg showed that the diffracted intensities destructive interference between waves scattered by a periodic from crystals of simple structures such as the alkali halides can array of objects; scattering by resonators was actually the subject be explained qualitatively in terms of scattering from point atoms of Ewald’s research at the time. Therefore, this single observation that have a scattering power proportional to their atomic numbers. When the atoms scatter in phase, their scattering power adds to- gether to give strong reflections, when they scatter in antiphase the * E-mail: [email protected] reflection is weak. This approach is sufficient to establish the basic

0003-004X/16/0005–1036$05.00/DOI: http://dx.doi.org/10.2138/am-2016-5473 1036

Downloaded from http://pubs.geoscienceworld.org/msa/ammin/article-pdf/101/5/1036/3603139/5_5473AngelCO.pdf by guest on 27 September 2021 ANGEL AND NESTOLA: A CENTURY OF MINERAL STRUCTURES 1037

architecture of simple minerals such as halite, and to determine the different atoms. It causes a general decrease in the structure factors, number of atoms in the unit cell of the structure (e.g., Bragg 1975). and hence diffracted intensities, with increasing diffraction angle When combined with the measured density and an accurate value (Debye 1914). Equation 2, which was later justified in terms of of Avogadro’s number (Millikan 1910), the unit-cell parameters the fundamental physics of X‑ray scattering by atoms (e.g., Waller of halides could be determined on an absolute basis. Subsequent and Hartree 1929), shows that the observed diffracted intensities application of what we now know as Bragg’s law allowed the represent a time-and-space average over all of the atoms within determination of X‑ray wavelengths, and then the determination the crystal. This average structure includes the static displacements of lattice parameters of new crystals without knowledge of their due to atoms occupying slightly different positions in different unit structures or densities. cells and the dynamic displacements due to temperature-induced The subsequent early development of diffraction crystallogra- vibrations of all of the atoms. In the 1920s the effects of atomic phy was remarkably rapid, both in experimental techniques and displacements on the diffracted intensities were subsumed into the in the theoretical developments necessary to determine structures scattering factors of atoms derived from intensity measurements of from the measured diffracted intensities. It was quickly recognized known structures, but nowadays the displacements are described that the idea of a simple reflection event by the lattice planes in by a model whose parameters are refined to the intensity data the crystal was inadequate to explain the diffracted intensities in along with the fractional coordinates of the atoms. detail. Darwin (1914) solved the problem of dynamical diffraction Thus, by 1920, all of the essential concepts of diffraction of X‑rays and thus identified the role of extinction, absorption, crystallography had been established and could be used to deter- and multiple diffraction in determining the intensities of diffracted mine the structures of crystals. They remain the basis of structure beams and established the concept of the mosaic crystal. This determination and refinement from X‑ray diffraction data to the concept, confirmed by experimental measurements of intensities, present day. The key steps in the determination of a crystal structure is that even gem quality crystals are actually comprised of small start with the collection of diffraction data, followed by “indexing” perfect crystal domains misaligned with respect to one another. (the consistent assignment of Miller indices to each diffraction We now understand that these domains are created by the defects peak). This allows the calculation of the unit-cell parameters from in the crystal structure, such as dislocation walls. The key point for the diffraction angles of each reflection, and the determination experimental measurements of X‑ray intensities is that the indi- of the space group symmetry from the pattern of systematically vidual domains do not diffract at exactly the same position because absent reflections. The measured intensities are then corrected for they are rotated with respect to one another, and this necessitates experimental factors and are reduced to the squares of structure the measurement of diffracted beam intensities by rotating the factors F2(hkl). This is not sufficient to determine the structure crystal in the beam during diffraction measurements (Bragg and because while the intensities are known, the phase factors a(hkl) Bragg 1913). It is the resulting integrated intensities (not the peak required to calculate the scattering density through Equation 1 are intensities), corrected for absorption and extinction and geometric not measured and are therefore unknown. Structure solution, or factors that depend on the geometry of the measurement, that are the determination of at least some atom positions within the unit related to the scattering density in the crystal (e.g., as summarized cell, is thus the biggest challenge in diffraction crystallography. in Bragg and West 1928). Bragg (1915a) expressed the relationship Once some atoms are located, phases a(hkl) can be estimated from between the absolute intensities corrected for experimental effects them using Equation 2. A Fourier synthesis (Eq. 1) with these and the scattering density in terms equivalent to a Fourier series, phases and the measured intensities is then used to determine the later (Duane 1925; Bragg 1929) written as: positions of any missing atoms. The test of a model structure is made through Equation 2, by comparing the values of the structure ρ = ⎡ π + + +α ⎤ (1) factors calculated from the structure model with those derived (xyz) ∑ F(hkl) ⎣⎢cos2 (hx ky lz) (hkl)⎦⎥ from the measured intensities. The subsequent step of structure The quantity |F(hkl)| is equal to the square-root of the absolute refinement is the adjustment of the structure model by adjusting intensity of the reflection hkl. It is the modulus (amplitude) of the the values of the atom coordinates and displacement parameters structure factor. The a(hkl) is the phase of the reflection. If the scat- to improve the fit to the measured intensities. tering density corresponds only to independent spherical atoms, The very first crystal structures to be determined had atoms in

each with a scattering power fi that is isotropic but depends on positions within the unit cells that were fixed by symmetry. The sin(q)/l, the Fourier series can be inverted to obtain the structure diffracted intensities could therefore be used to calculate the actual

factor equation: scattering factors fi of the individual atoms, which could then be used in the analysis of more complex structures. Later, scattering F(hkl) = f ⎡cos2π hx + ky + lz + isin2π hx + ky + lz ⎤ B hkl factors were determined by direct calculation from a model of the ∑ i ⎣⎢ ( i i i ) ( i i i )⎦⎥ i ( ) i=1,n electronic structure of isolated atoms or ions (e.g., Brown et al. (2) 2006). By the late 1920s it is apparent from papers (e.g., Bragg This equation expresses the value of the structure factor of and West 1928) that the methods of data collection and reduction any reflection in terms of a summation of the scattering power of had become routine; Brown and West (1927) describe their mea- the atoms within one unit cell, with phase factors that depend on surements as “The usual X‑ray rotation photographs...” Structure the atom positions expressed in fractional coordinates x, y, z and solution was the challenge. It proceeded by a mixture of deduction the Miller indices hkl of the reflection. The effects of local atom from the chemical analysis combined with determination of the displacements on the structure factors are expressed by the factor space group symmetry of a crystal, and by use of the principles

Bi, which in general is an anisotropic function that is different for of crystal chemistry as they emerged from previous structure

Downloaded from http://pubs.geoscienceworld.org/msa/ammin/article-pdf/101/5/1036/3603139/5_5473AngelCO.pdf by guest on 27 September 2021 1038 ANGEL AND NESTOLA: A CENTURY OF MINERAL STRUCTURES

solutions. Bragg (1937, page 27) characterized early structure exactitude with which atomic coordinates can be determined is solution as “guessing the answer to a puzzle.” The idea that the only limited by the amount of labor expended in making measure- O atoms in structures often approximated close-packed layers ments”; the experience of his assistant Taylor (1933) shows that quickly emerged from studies of (Bragg 1915b) and simple once the basic crystal structure was solved, structure refinement oxides, sulfides, and halides (Bragg 1920) and it was found to be was clearly limited not by quantity or quality of data, but by lack true also for olivine-type structures (Bragg and Brown 1926a). of computing power. The final individual cation- (T-O) This allowed structure solution to proceed by considering how bond lengths in the tetrahedra of Taylor’s initial structure to fit the close-packed arrays into the unit cell and then deciding ranged from 1.55 to 1.75 Å, with the conclusion drawn that “both which interstices in the oxygen array were occupied by cations tetrahedral groups are nearly regular” (Taylor 1933). These state- (Bragg and Brown 1926b). The crystal structure of and its ments indicate that uncertainties in cation-oxygen distances were high-temperature polymorphs showed that Si uniformly preferred of the order of 0.1 Å at best, quite insufficient to draw conclusions tetrahedral coordination in common minerals. Other structure de- about the detailed crystal-chemistry, to follow structural changes terminations showed that Al and Si could substitute for one another due to changes in P, T, or phase transitions, or even to determine in tetrahedra, that Al could also be octahedrally coordinated, while basic patterns of Al,Si ordering. As feldspars exhibit both dis- cations such as Fe and Mg normally occupied octahedral sites. It placive transitions and various convergent and non-convergent should be emphasized that such crystal-chemical principles that ordering processes that greatly influence their thermodynamic were codified as “laws” by Goldschmidt (1926) following Bragg’s properties, the development of methodologies in diffraction crys- (1920) ideas about ionic radii, translated for American Mineralo- tallography as they were enabled by increased computing power gist by Wherry (1927) and expressed in terms of the “coordina- can be followed by seeing how knowledge of the struc- tion method” (Pauling 1929) or polyhedral model, were used tures became more and more precise. The basic theme here is that as hypotheses to obtain initial structure solutions. The structure improved precision in structure refinements means more precise solutions, and hence the crystal-chemical ideas, were then tested determination of the atomic coordinates, which in turn requires by comparison to the measured intensities through the structure more precise and higher-resolution data. More precise data comes factor equation. There were still significant potential pitfalls in from both better measurement methods of the raw diffraction data structure determination, as can be seen in the first paper to report and more precise corrections for instrumental and experimental a full structure determination in American Mineralogist (Buerger effects on the measured diffracted intensities. Higher-resolution 1931). In determining the structure of marcasite Buerger (1931) data also implies more data. Therefore, crystallographers needed showed that two previous determinations by diffraction were in more computing power to obtain more precise crystal structures. error because of different errors in the author’s setting of the space Attempts in the 1930s to provide aids for structure refinement group symmetry, while a third proposal although correct, was “of were restricted to graphical tools such as structure-factor graphs the nature of guesswork.” (Bragg and Lipson 1936) and mechanical tools such as Beevers- By 1930 (e.g., Gruner 1929, 1931; Bragg 1930) the basic Lipson strips (Beevers and Lipson 1934) both of which eased the architecture of the crystal structures of most major rock-forming calculation of the terms in Fourier series such as Equations 1 and minerals and large numbers of sulfides, elements, and oxides had 2. They enabled Chao et al. (1940) and Cole et al. (1949) to refine been established by X‑ray diffraction. The framework structures the crystal structure of an , starting from the sanidine of zeolites and feldspars were the major exceptions because their model of Taylor (1933). Although the method of least-squares low packing density meant that ideas of close-packed layers are not refinement was first applied to a crystal structure refinement by appropriate to describe these structures (Pauling 1929; Taylor et Hughes (1941), and the first feldspar structure to be refined this al. 1934). Thus initial models for structure solution were not avail- way was reported by Sörum (1953), the calculations were done able. Matchatschki (1928) argued from chemical considerations by hand, severely limiting the number of the data that could be and knowledge of established alumino-silicate crystal chemistry handled. Nonetheless, Sörum (1953) was able to demonstrate from

that feldspars must be a fully connected network of AlO4 and SiO4 differences in tetrahedral bond lengths that the ordering pattern tetrahedra. Taylor (1933) solved the structure of sanidine using in anorthite-rich intermediate is different from that this idea and knowing the cell dimensions, symmetry and, criti- in . The introduction of mechanical calculators (“Hollerith

cally, the size of (Si,Al)O4 tetrahedra from other structures. On machines”) in the late 1940s and 1950s enabled larger data sets to Christmas Day 1932 (see Smith 1980) he solved the geometric be handled, with attendant increase in precision. The architecture problem of how to fit the necessary four tetrahedra as a ring into of mechanical calculators made them more suitable for Fourier the asymmetric unit of the unit cell, without distorting them, by calculations (Cox and Jeffery 1949; Hodgson et al. 1949) than for having the insight to place the external oxygen atoms of the ring the implementation of least-squares refinements even though it was on symmetry elements so that the connectivity of the framework recognized this was not the best way to determine accurate bond was generated naturally by the space group symmetry. The K lengths and angles (Cochran 1948). So refinement of structures atoms were then located by comparing observed and calculated continued to use the step-wise cycling of structure factor calcula- intensities of a few reflections, and the final refinement of the tions from trial coordinates (Eq. 2), which were used in Fourier atomic positions proceeded, as for other structures determined in syntheses of the electron density through Equation 1, from which that era by “a laborious search for more accurate atomic parameters better trial coordinates could be determined. Automation with Hol- which will give calculated F-values in agreement with the observed lerith machines, coupled with the assistance of several colleagues, relative intensities of a large number of reflections” (Taylor 1933). wives, and students to feed them with punched cards did allow Thus Bragg (1930) was wrong when he had suggested that “The the number of data used in a refinement to be rapidly increased.

Downloaded from http://pubs.geoscienceworld.org/msa/ammin/article-pdf/101/5/1036/3603139/5_5473AngelCO.pdf by guest on 27 September 2021 ANGEL AND NESTOLA: A CENTURY OF MINERAL STRUCTURES 1039

Bailey and Taylor (1955) used 2650 reflections (and four assistants diffractometers to collect the data, resulted in the publication of including Mrs. Bailey) to show from the T-O bond lengths that the about 150 feldspar structures over the two decades 1962–1982. structure of was effectively fully ordered with respect The improvements in diffractometry and data processing led to to Al and Si distribution. T-O bond lengths being determined to a reproducibility of ~0.002 The year 1962, a half-century from the discovery of X‑ray Å (e.g., Dal Negro et al. 1978), enabling the details of the crystal diffraction, divides the era of “discovery” of the basic architec- architecture to be determined and how the T-O bond lengths change ture and principles of mineral structures from the modern era of with the Al,Si content of the tetrahedra (e.g., Brown et al. 1969; investigating the detailed crystal chemistry of minerals. By 1962 Kroll 1983; Kroll and Ribbe 1983; Ribbe 1994). there had been 11 full structure refinements of feldspars (Ribbe By the late 1980s computing power was no longer a limita- 1994), made over a period of 30 yr since the original structure tion to structure refinements of minerals; essentially arbitrarily determination of sanidine by Taylor (1933). With the larger data large data sets could be used for full refinements of not only sets collected in the 1950s and processed with Hollerith machines the positional parameters of the atoms, but also the anisotropic and early mainframe computers, the refined structures had a displacement parameters to describe the thermal vibrations of the claimed precision in bond lengths of 0.002 Å (e.g., Cole et al. atoms, without which the positions of the atoms can be biased to 1949) although this uncertainty is clearly underestimated because incorrect values (see Eq. 2). The precision of refined structures it is derived directly from the difference Fourier maps (Cochran was therefore limited by the use of point detector diffractometers. 1948; Cruickshank 1949a, 1949b). A more realistic estimate from To achieve higher resolution many more reflections at higher the least-squares refinement by Sörum (1953) would be of the diffraction angles have to be collected that, with a point detector order of 0.005–0.01 Å, which is sufficient to determine the basic diffractometer, means a very large increase in data collection times pattern of Al,Si ordering from the difference between tetrahedra to the order of weeks because the data are collected one reflection occupied only by Al (Al-O ~1.74 Å) and Si (Si-O ~1.62 Å) but at a time. For example, the extension of a data collection on albite not really good enough to determine intermediate states of order, from 60° to 80°, which reduces the formal estimated standard such as that found in orthoclase (Cole et al. 1949; Jones and Taylor deviations (e.s.d. values) on T-O bond lengths to 0.001 Å (e.g., 1961). Fourier maps showed that the extra-framework cations in Meneghinello et al. 1999) requires the measurement of about albite, at least, exhibited anisotropic apparent motion (Ferguson twice the number of reflections, with the additional higher-angle et al. 1958), but anisotropic displacement parameters had not been reflections being systematically much weaker and thus requiring refined for them or the framework cations. much longer to collect to the necessary precision. Although structures based on intensities collected by photo- In the anorthite-rich plagioclase feldspars where the ordering graphic methods continued to be reported well into the 1960s, of Al and Si (and accompanying structural changes) gives rise to the first structure refinement of a feldspar of the modern era was systematically weak reflections the challenge is greater because published by Onorato et al. (1963). They collected the intensity these weak reflections have to be collected in sufficient numbers data not by film, but with a commercial diffractometer, equipped with sufficient precision (which means longer counting times) to with a point detector that recorded the intensity data digitally and determine precise bond lengths within the structure. Even with a Mo X‑ray tube that, because of the shorter wavelength, makes rotating anode sources that became available for laboratories in absorption less of a problem than with Cu radiation. Onorato et the late 1980s a data collection for such a plagioclase would still

al. (1963) performed a structure refinement using the full-matrix require several days to collect ~3300 reflections to 2qmax = 60° (e.g., least-squares method on a computer. The availability of University Angel et al. 1990). The introduction of CCD-based area detectors mainframe computers at the beginning of the 1960s that could run for diffractometry speeded up data collections by allowing many programs written in high-level languages (mostly Fortran) that reflections to be collected simultaneously. Subsequent improve- were compiled locally (making them portable) opened the mod- ments in detector technology, which have greatly reduced the ern era of diffraction crystallography. Programs were developed intrinsic noise in X‑ray detectors while increasing their dynamic for data reduction including face-based absorption corrections range, have allowed data collection times to be further greatly (e.g., Busing and Levy 1957; Burnham 1966) to provide more reduced. accurate data, and programs to perform full-matrix least-squares As an example of modern laboratory capabilities, Figure 1 il- refinement of the structure directly to the structure factors via lustrates the uncertainties in T-O bond lengths in end-member Val Equation 2 became available (e.g., Busing et al. 1962; Stewart et Pasmeda anorthite (Carpenter et al. 1985) determined from a series al. 1972; Finger and Prince 1974; Sheldrick 2008). As computer of data collections. We measured a crystal of ~70 mm diameter with storage and power increased, structures were refined to more and a commercial instrument, the Oxford Diffraction SuperNova dif- better data, which enabled a better description of the scattering fractometer equipped with a Mo-target X‑ray micro-source and a density in terms of the anisotropic displacement parameters of Pilatus 200 K area detector1. The detector has no intrinsic noise and the atoms. Thus while Onorato et al. (1963) reported a full-matrix has no read-out delay, further speeding data collections compared least-squares refinement of sanidine with individual isotropic to CCD detectors. The ability to collect data faster also means that displacement parameters for each atom (a total of 28 parameters) 1 Deposit item AM-16-55473, CIFs. Details of the data collections, data reduction, to 960 reflections, within a few years it was practical to perform and structure refinement procedures, and the final refined crystal structures, are fully anisotropic refinements such as that of albite (Wainwright and provided in the CIF deposited as supplementary material. All structure refinements Starkey 1967) with 117 parameters refined to 3530 data measured were performed with RFINE-2015, a development version of RFINE4 (Finger and Prince 1974). Deposit items are free to all readers and found on the MSA web on a counter diffractometer. The increase in computing power, site, via the specific issue’s Table of Contents (go to http://www.minsocam.org/ together with the widespread introduction of computer-controlled MSA/AmMin/TOC/).

Downloaded from http://pubs.geoscienceworld.org/msa/ammin/article-pdf/101/5/1036/3603139/5_5473AngelCO.pdf by guest on 27 September 2021 1040 ANGEL AND NESTOLA: A CENTURY OF MINERAL STRUCTURES

multiple measurements can be made of the same, or symmetry- refinement models and refinement parameters such as weighting equivalent, reflections from the crystal at different positions of the schemes (Fig. 2). In this sense, the determination by modern X‑ray diffractometer. This redundancy in the data allows, with modern diffraction of the geometric details of the average crystal structure computing power, the intensities to be corrected more accurately in terms of atom positions and interatomic distances and angles for experimental factors such as absorption by the crystal and its is reliable and robust. mount, the variations in incident beam intensity, and the detector response. As a consequence, as an extreme example, a partial data Beyond geometry: Chemistry and collection of just 90 s on anorthite is sufficient to determine the thermodynamics of minerals bond lengths to an accuracy of 0.02 Å (Fig. 1) and thus the es- Because accurate X‑ray diffraction intensity data can be col-

sential architecture of the structure! A full data set to 2qmax = 60° lected precisely and quickly with high redundancy, we now focus collected in 48 min yielded e.s.d.(T-O) = 0.004 Å, equivalent to on the use and applicability of structure refinements to solving the results from 6 days data collection 25 yr ago from a signifi- mineralogical problems beyond that of simply reporting the cantly larger crystal (Angel et al. 1990) or several months in the structural architecture, bond lengths and angles. Chemical and

1950s (Kempster et al. 1962)! A data collection to 2qmax = 100° thermodynamic information about minerals can be obtained by yielded in excess of 95 000 individual reflections, took 81 h and determining the occupancies by different atomic species of indi- gives e.s.d.(T-O) = 0.0005 Å. The refined structures of anorthite vidual crystallographic sites within the mineral. Cation partitioning show no significant differences in bond lengths when the data are between sites within a single mineral crystal has great potential

restricted to lower resolution (down to 2qmax = 50° for MoKa), or to be used as a single-phase geo-thermometer or even a geo- with decreasing exposure time at constant resolution in full data speedometer, especially because the process of re-equilibration sets (Fig. 1). The small offsets of bond lengths when the data are of the site occupancies after temperature change is rapid as the

restricted to very low resolution (2qmax < 50°) arises from correla- distances over which cations have to be exchanged is only of the tions between the positional and displacement parameters in the order of the size of the unit cell. This contrasts with chemical refinement that are only suppressed with higher-resolution data. re-equilibration between different mineral phases that require The e.s.d. values in bond lengths for full data sets scale approxi- diffusion over distances that are of the grain scale. Determination mately with the inverse square root of the number of reflections of the site occupancies also, obviously, allows the composition of used in the refinement, which means that modern diffraction data the mineral to be obtained. Thus diffraction is an important tool such as these are precise and accurate. Furthermore, when the data for determining compositions of minerals that cannot be analyzed are collected with sufficient resolution, conventional refinements by microprobe, for example inclusions trapped within diamonds also yield bond lengths that within the e.s.d. values are mostly (e.g., Nestola et al. 2011) or other host minerals where extraction independent of experimental influences and detailed choices of of the inclusion mineral is not permitted because it requires at least partial destruction of the host mineral. Chemical site occupancies can be obtained by direct structure refinement of the occupancies, or indirectly from the bond lengths, and we will discuss one ex- ample of each of these in turn. In doing so we will illustrate that the limitations of modern X‑ray structure refinement are not in the diffraction data, but in the nature of the refinement models that are used to interpret the diffraction data and the crystal-chemical

Figure 1. The precision of X‑ray structure refinements as illustrated Figure 2. There is no significant variation in individual -T O bond by average bond lengths of the Si-rich and Al-rich tetrahedra in anorthite lengths in anorthite with different refinement models, as illustrated for (open symbols) and one individual Al-rich and one Si-rich tetrahedron the T1oo tetrahedron. Refinements are (1) using robust-resistant weights (closed symbols with error bars). Data from Kempster et al. (1962), and refinement to structure factors F, (2) unit weights and refinement to Wainwright and Starkey (1971), Angel et al. (1990), and the current F, and (3) normal weights and refinement toF 2. The intensity data set had work. The deviations in earlier work are the consequence of limited data a resolution corresponding to 60° 2q for MoKa radiation (labeled “60” in resolution. Fig. 1). E.s.d. values are approximately equal to the symbol size.

Downloaded from http://pubs.geoscienceworld.org/msa/ammin/article-pdf/101/5/1036/3603139/5_5473AngelCO.pdf by guest on 27 September 2021 ANGEL AND NESTOLA: A CENTURY OF MINERAL STRUCTURES 1041

models used to interpret them. Angel et al. 1990; Meneghinello et al. 1999). When the contrast in X‑ray scattering powers of the atoms of When the contrast in scattering power between the atoms is interest is small, for example as between Al and Si, the site oc- significantly greater than that between Al and Si, it is possible to cupancies cannot be determined by direct refinement. However, determine the site occupancies by refining the scattering power the example of the anorthite structure discussed above shows that of each atomic site. This can be done in two ways. When just two the distances between the atom centers can be determined to high atomic species are known to occupy a given site, and the site is precision by conventional structure refinement to modern diffrac- assumed to be fully occupied, then the occupancies can be refined tion data (Figs. 1 and 2). If we suppose that T-O bond lengths in a directly. An alternative is to refine the occupancy of an atom chosen feldspar vary linearly with Al content and that the difference be- to represent the average scattering curve of the multiple occupants tween Al-O and Si-O is ~0.12 Å (e.g., Kroll and Ribbe 1983) then of the site, and then the actual site occupancies can be inferred from an uncertainty of the order of 0.0005 Å in a T-O bond length (Fig. the total refined scattering power (measured in electrons) and the 1) should allow Al contents to be determined with an uncertainty atomic numbers of potential occupants. Clearly, because structure of less than 1%. However, while the primary factor controlling refinement only yields the total scattering of the site through the individual T-O bond lengths is the Al,Si content of the site, several scattering factor f (Eq. 2), the interpretation of the result in terms secondary factors have also been shown to be significant. These of multiple occupancies requires further constraints or assumptions include the occupancy of the adjacent tetrahedra, the bonding of about the chemistry and the distribution of species between sites. the extra-framework cation to the oxygen, and the T-O-T and O- One obvious additional constraint that can be used for structures at T-O angles (e.g., Phillips and Ribbe 1973; Phillips et al. 1973). ambient pressure and temperature are the bond lengths of the site, This explains why the individual bond lengths in any one of the if the atomic species of interest give rise to significantly different

nominally pure SiO4 tetrahedra in anorthite range from <1.59 to bond lengths. Such a combined approach has been successfully ~1.65 Å (e.g., Fig. 2). This variation is more than half the expected used in several mineral systems such as spinels (Della Giusta et

difference between the bond lengths for pure AlO4 and SiO4 tetra- al. 1996; Lavina et al. 2002). hedra! Some of the geometric effects appear to be compensated In mantle a simplified intermediate approach can internally within an individual tetrahedron, so it was recognized be used because it is known from chemical analyses of olivine (e.g., Kroll 1983; Kroll and Ribbe 1983) that mean bond lengths inclusions in diamonds that their compositional variation is of tetrahedra were more reliable indicators of Al,Si content dominated by variation in Mg,Fe content (Stachel and Har- than individual bond lengths. As more precise structures became ris 2008). These atoms have sufficiently different scattering available it has become apparent that the mean values in powers that direct refinement of the atomic occupancies on plagioclase feldspars are sensitive to Ca,Na content, and can be the two M sites within the structure is possible, subject to the perturbed by different refinement models for the partially disor- sites being assumed to be fully occupied. Mantle olivines also dered structures (Angel et al. 1990; Ribbe 1994). There are further typically contain trace amounts of other elements, mostly Ca, difficulties. It is not possible to determine by experiment whether Mn, and Ni. The scattering powers of Ni and Mn for X‑rays anorthite and albite are fully ordered; techniques such as neutron are very similar to that of Fe, so their contribution to the X‑ray diffraction and NMR have uncertainties of a few percent in the diffraction intensities is effectively included in the refined Fe determination of Al and Si site occupancies. Therefore we do not content. As Ca has an atomic number approximately halfway

know the true value of in pure AlO4 or pure SiO4 tetrahedra. between that of Mg and Fe, it can be considered as contributing Neither do we know whether the differences in values for equally to the refined Mg and Fe contents. We will denote the the individual Si-rich tetrahedra, for example as shown in Figure Mg occupancy obtained by structure refinement using only Mg 1, represent small differences in Si,Al contents or are simply the and Fe as the site contents as Mg*. Its value should be equal to effects of other factors on bond lengths. The refined displacement the fraction (Mg+0.5Ca)/(Mg+Fe+Mn+Ni+Ca) determined by parameters of the atoms also correspond to apparent displacements chemical analysis. Figure 3 shows the site occupancies refined of T and O atoms along the bond of the order of 0.07–0.08 Å, but to a single data set collected from a mantle olivine crystal of X‑ray Bragg diffraction as expressed by Equation 2 does not tell approximate size 150 × 110 × 80 mm, out to a resolution limit us how these displacements are correlated. Different models for of 100° 2q with an average redundancy (number of symmetry- interpreting the refined displacement parameters in terms of cor- equivalent reflections measured) of 6, in 6 h. The composition related, uncorrelated or riding motions of one atom with respect has been measured by chemical analysis as Mg* = 0.877(1). It to the other (e.g., Finger and Prince 1974) yield corrections to T-O is obvious that the refinement performed with scattering factors bond lengths that are of the order of 0.02 Å. All of these factors calculated for neutral atoms (the default for refinements of mo- mean that the absolute occupancy of tetrahedral sites in feldspars, lecular structures in programs such as SHELX, Sheldrick 2008) or similar alumino-silicates such as zeolites, cannot be established yields an overall composition from the refined occupancies of to the precision that would be expected from the precision in bond the two M sites that is about 2–3% too high in Mg. The actual lengths. However, all of these limitations arise strictly from the value also changes by >1% in Mg* depending on the resolution refinement models and crystal-chemical analysis. This means of the data used. The agreement with the known composition that if different crystal structures of the same type (e.g., albites) is better if scattering factor curves for charged atoms are used are refined with exactly the same model, then X‑ray structure (Fig. 3a), but there is still significant variation in the results refinement can yield precise values for the differences between with data resolution. Furthermore, the values and the variation the structures, and thus changes in the state of Al,Si order, even if with resolution of Mg* on individual sites and the overall com- the absolute values of site occupancies are less well-defined (e.g., position is different if different weighting schemes and/or the

Downloaded from http://pubs.geoscienceworld.org/msa/ammin/article-pdf/101/5/1036/3603139/5_5473AngelCO.pdf by guest on 27 September 2021 1042 ANGEL AND NESTOLA: A CENTURY OF MINERAL STRUCTURES

criteria for rejecting outlier data are changed (compare Fig. 3b in the refinement model. The conventional refinement model that with 3a). If corrections for secondary extinction (the reduction we have discussed so far relies on scattering factors for atoms in measured intensities due to interactions between the incident that are calculated for spherical isolated atoms. For this reason it and diffracted beams in the crystal, Becker and Coppens 1974) is sometimes called the “independent atom model” (IAM). But are not made, then the value of Mg* can be shifted by 0.2%. the measured intensities from crystals are scattered by all of the Therefore, because of the sensitivity of the results to refinement electrons, including those involved in bonding. Kroll et al. (1997) conditions, conventional refinements using charged atom form showed that the variation in refined site occupancies of orthopy- factors cannot yield the composition of a mantle olivine more roxenes with 2q range of the data arises from the change in the reliably than ±1%. contribution of the bonding electrons to the observed intensities In the past some of this variation in refined site occupancies as 2q changes. In simple terms, at low Bragg angles all electrons has been attributed to short-comings in the experimental correc- including the bonding electrons contribute to the diffracted tions to the data as well as to different choices of scattering factors intensities. At high angles the broader spatial distribution of the as we have discussed above (Kirfel 1996). However, refinements bonding electrons means that they contribute disproportionately to structure factors calculated by Fourier transformation of the less to the scattered intensities, and the measured X‑ray intensi-

electron density of end-member Mg2SiO4 determined by DFT ties are therefore dominated by scattering from the core electrons computer simulation also show a strong variation in refined site of each atom. That is also why charged-atom form factors give occupancies with resolution (Fig. 4a). This indicates that the the correct Mg* value for olivine when we use high-angle data, experimental data are not at fault, but that there are short-comings while neutral-atom form factors yields Mg* values that are ~0.02 too high (Fig. 3). The variations in Mg* with data resolution arise from correlations between the occupancies, the scale factors, the extinction parameters, and the parameters used to describe the displacements of the atoms due to thermal and static positional disorder. As for the bond lengths alone (Fig. 1) these correlations are only suppressed by the use of both high- and low-resolution

Figure 3. The results of site occupancy refinements of a natural olivine with composition Mg* = 0.877(1) as a function of data resolution. (a) Refinements of a conventional independent atom model (IAM) with neutral-atom scattering factors yield total Mg* values that are 2–3% too high. Charged atom refinements give results that are correct to within Figure 4. The results of site occupancy refinements to a synthetic 0.01 in Mg*. (b) The values and trends with resolution of Mg* from data set calculated from a DFT simulation of end-member forsterite. (a) a conventional IAM refinement change are sensitive to the weighting Refinements of a conventional independent atom model (IAM) with scheme; the results from a robust-resistant refinement are shown. (c) charged-atom scattering factors yield total Mg* values that are too low, and Refinement of the electron-in-bond model of Heinemann et al. (1999) that vary significantly with data resolution. (b) Refinement of the electron- yields Mg* within 2 e.s.d. values of the known composition, independent in-bond model of Heinemann et al. (1999) yields Mg* within 2 e.s.d. of resolution. The darker band is ±1 e.s.d., the lighter band ±2 e.s.d. values. values of the correct value (1.0), with less variation with data resolution.

Downloaded from http://pubs.geoscienceworld.org/msa/ammin/article-pdf/101/5/1036/3603139/5_5473AngelCO.pdf by guest on 27 September 2021 ANGEL AND NESTOLA: A CENTURY OF MINERAL STRUCTURES 1043

intensity data, for which the displacement and occupancy factors time scales are shorter, or alternatively the crystal sizes can be have significantly different contributions (Eq. 2). smaller. At high pressures, high temperatures, or in operando, the Kroll et al. (1997) suggested that the problem could be ad- data quality remains in the majority of cases the factor limiting dressed by removing the correlation between the site occupan- the accuracy and precision of the resulting structures. Such data cies and the displacement parameters simply by fixing the latter normally require additional corrections for absorption by the at values that ensure that the refined site occupancies became environmental cell but, at the same time, the cell often restricts independent of data resolution. This procedure was somewhat both the resolution and redundancy of the data, making such cor- artificial and also very time consuming. Another approach was rections more difficult to perform. The real challenge is therefore to refine the effective charges on the atoms, by refining a mixture to make the data corrections accurately based on fewer data. For of fully charged and neutral atoms on each site in the structure, laboratory high-pressure studies the lack of redundant data means a careful study of several stoichiometric end-member minerals that the best approach is to base the corrections on appropriate (Sasaki et al. 1980) having indicated that the scattering density physical models of both the crystal and the pressure cell. But was best represented by partially charged atoms subject to the the data quality never matches what can be achieved with the constraint that the resulting model is charge neutral. This result same instrument from the same crystal in air. An alternative is was applied (e.g., Rossi et al. 1983; Hawthorne et al. 1995) to provided by recent developments in high-pressure single-crystal determine site occupancies by site occupancy refinement using diffraction on synchrotron X‑ray sources, where the use of much fixed partial charge scattering factors. While releasing the con- higher-energy, much more intense and much smaller X‑ray straint of charge balance can sometimes provide a better fit to beams reduces the magnitude of the experimental corrections the diffraction data, it can result in a structure model that has a required while at the same time providing raw intensity data net charge and cannot be considered physically realistic. of greater reciprocal space coverage, greater redundancy and Fourier maps of the calculated electron density (Eq. 1) from much improved signal to noise (e.g., Hejny and Minkov 2015). conventional refinements of silicates indicate that the major As a consequence, the resulting quality of crystal structures at misfit of the density is approximately midway along the Si-O pressures of tens of GPa is starting to approach that which can bond. This confirms that conventional refinements do not model be determined at room pressure which opens a new frontier in the contribution of the bonding electrons (e.g., Hawthorne et al. ultrahigh-pressure crystal chemistry (e.g., Finkelstein et al. 2014; 1995). Heinemann et al. (1999) therefore implemented the idea of Glazyrina et al. 2014). Kirfel and Will (1980) to modify the crystal structure model in the Laboratory X‑ray diffraction data from crystals in air is now refinement in a manner that directly reflects this short-coming of of the quality where the results are limited not by the data but by a conventional independent atom model. Heinemann et al. (1999) the physical validity of the refinement models used to interpret showed that by refining a structure model that has one electron the measured data. Thus the standard structural models that are placed on the Si-O bonds (and then refining position along the conventionally used in refinements are based on independent bond and the displacement parameters of the electron), the con- atomic spherical scatterers and thus can be said to “neglect the tribution of the bonding electrons to the diffracted intensities can chemistry”—they do not model the modification of the electron be successfully modeled. In a simplification of this model we density of the atom due to bonding. For most minerals, this find that fixing the electron to a point 70% along the Si-O bond modification of the electron density is very small, which is why toward O, and restricting all such electrons in the refinement the structural geometry from refinement of an IAM is accurate model to have the same isotropic displacement parameters is to a level limited by the intrinsic noise of the data. We have seen sufficient to eliminate nearly all of the variation in site occupan- that the site occupancies are more affected and that the electron-

cies with 2q range (Fig. 4b) for the DFT data set of Mg2SiO4. in-bond model (Kirfel and Will 1980; Heinemann et al. 1999) is Refinement of such a model with robust-resistant weights and a a simple but under-utilized extension of the IAM, which can be realistic extinction model (Becker and Coppens 1974) to the data readily implemented at no intellectual difficulty, tested, and used of the natural olivine results in Mg,Fe site occupancies that are to provide site occupancies with a reproducibility that is better in agreement with the composition determined by microprobe than a routine microprobe analysis for simple chemical systems analysis. The accuracy of these X‑ray structure refinements (Fig. with high contrast in atomic number such as Fe,Mg in olivine. 3c), is as good as that achieved by chemical analysis. However, the determination of site occupancies in even simple systems to a precision better than 0.001 is unrealistic given the Implications short-comings in the IAM and the Bragg approach embedded What we have shown in this paper is that the precision with in Equation 2 (see also Heinemann et al. 1999; Stimpfl et al. which we can determine mineral structures has evolved over 1999; Merli et al. 2000). In particular, in a crystal structure with 100 yr in parallel with the developments in X‑ray diffraction mixed site occupancies such as Al,Si in feldspars or Mg,Fe in crystallography that have improved the quality of the diffraction olivines, the detailed local structural geometry has been shown data, and with the increased technological capacity to handle by various local experimental probes such as NMR and IR and the data. With modern instruments on laboratory sources Bragg DFT simulations to be dependent on the local arrangement diffraction data can be collected from small mineral crystals of and short-range ordering of the cations. Therefore, the oxygen 50–100 mm size at room conditions in a few hours that is of such “positions” obtained from structure refinement are some kind of quality that the accuracy and precision of the resulting structural average position of their true local positions, and that is why the parameters such as bond lengths and angles is beyond that re- refined structure exhibits bond lengths that reflect some kind of quired for crystal-chemical analysis. For synchrotron sources the average of the bond lengths in the pure end-member structures.

Downloaded from http://pubs.geoscienceworld.org/msa/ammin/article-pdf/101/5/1036/3603139/5_5473AngelCO.pdf by guest on 27 September 2021 1044 ANGEL AND NESTOLA: A CENTURY OF MINERAL STRUCTURES

At the same time, this average bond length is in some way also crystalline state, such as surfaces and nanoparticles, all of whose modified by the correlation of positions with the refined displace- intriguing properties can only be understood in the context of struc- ment parameters, as is particularly evident in plagioclase feldspars tural deviations from the ideal average structure of the bulk mineral. (Angel et al. 1990). With increased data quality, the IAM can also be extended by refining parameters that correspond to the explicit Acknowledgments population and deformation of the electron density of each atom. We thank Editor Keith Putirka for insisting that this contribution to the (ap- proximately) joint centennial of the MSA, the American Mineralogist, and X‑ray Such “pseudo-atom” (Stewart 1976) or multipole models have been diffraction should be our top priority. It afforded us the excuse and pleasure of very successful in yielding the details of not only the bonding den- reading the historical literature of mineralogical crystallography, to which we were sity in pure materials such as (Overgaard et al. 2014), but even patiently guided by the staff of the library of the Department of Geosciences of the University of Padua. Early education in the history of feldspars was provided the core electron deformations in diamond (Bindzus et al. 2014). many years ago by M.G. Bown and P. Gay, and in X‑ray structure refinement by However, multipole models require an explicit atomic basis and Charles Prewitt and Larry Finger. Mauro Prencipe performed the DFT simulations are therefore not meaningfully defined for mixed-occupancy sites and generated the structure factors of forsterite, and Sula Milani performed the data collections and microprobe analyses on the San Carlos olivine with the help of typical of mineral solid solutions. The iterative Hirshfeld method Daria Pasqual and Raul Carampin. We thank Gilberto Artioli for the loan of many is an intermediate approach in which the structure is refined with older crystallographic texts, and him, David Allan, Dan Frost, and Diego Gatta for discussions and suggestions. This work was supported by ERC starting grant multipole models, and then density-functional simulations are used 307322 that funds the INDIMEDEA project to Fabrizio Nestola. to re-calculate the scattering factors so as to include all of the elec- tron density in the effective scattering factors, which are then used References cited in the next cycle of refinement (Capelli et al. 2014; Farrugia 2014). Andersen, C.W., Bremholm, M., Vennestrøm, P.N.R., Blichfeld, A.B., Lundegaard, Data quantity and quality has enabled an alternative approach, L.F., and Iversen, B.B. (2014) Location of Cu2+ in CHA zeolite investigated by X‑ray diffraction using the Rietveld/maximum entropy method. IUCrJ, 1, 382–386. known as the maximum entropy method (MEM), whose essence Angel, R.J., Carpenter, M.A., and Finger, L.W. (1990) Structural variation associated is to refine with the minimum constraints of “physical reasonable- with compositional variation and order-disorder behavior in anorthite-rich feldspars. ness” the electron density of the crystal simply by adjusting the elec- American Mineralogist, 75, 150–162. Bailey, S.W., and Taylor, W.H. (1955) The structure of a triclinic felspar. Acta tron density in a pixelated array of boxes across the unit cell. While Crystallographica, 8, 621–632. it still only provides an electron density averaged over all cells of Becker, P., and Coppens, P. (1974) Extinction within the limit of validity of the Darwin transfer equations. I. General formalisms for primary and secondary extinction the crystal, MEM results are model independent, which provides the and their application to spherical crystals. Acta Crystallographica, A30, 129–147. possibility of visually and directly locating non-framework species, Beevers, C.A., and Lipson, H. (1934) A rapid method for the summation of a two-di- such as the cations in zeolites, based only on observed structure mensional Fourier series. Philosophical Magazine, 17, 855–859. Bindzus, N., Straasø, T., Wahlberg, N., Becker, J., Bjerg, L., Lock, N., Dippeld, A.-C., factors (Andersen et al. 2014). So MEM gives electron densities and Iversen, B.B. (2014) Experimental determination of core electron deformation unbiased by a structural model but, like the IAM and its extensions, in diamond. Acta Crystallographica, A70, 39–48. the interpretation of this scattering density requires assumptions Boffa-Ballaran, T., Frost, D.J., Miyajima, N., and Heidelbach, F. (2010) The structure of a super-aluminous version of the dense hydrous-magnesium silicate phase D. about oxidation states, local ordering and distortions in structures American Mineralogist, 95, 1113–1116. with mixed site occupancies. We can get accurate electron densities Bragg, W.H. (1915a) Bakerian Lecture—X‑rays and crystal structure. Philosophical Transactions A, 215, 253–274. but not occupancies from X‑ray diffraction measurements, which ——— (1915b) The structure of magnetite and the spinels. Nature, 95, 561. only serves to illustrate the fundamental point that while an X‑ray Bragg, W.L. (1920) The arrangement of atoms in crystals. Philosophical Magazine, diffraction experiment, properly performed, provides an unbiased 40, 169–189. ——— (1929) The determination of parameters in crystal structures by means of Fourier map of the average electron density of a crystal, the interpretation series. Proceedings of the Royal Society of London A, 123, 537–559. of the electron density is model dependent. ——— (1930) The structure of silicates. Zeitschrift für Kristallographie, 74, 237–305. In the past, tour-de-force solution and refinement of crystal ——— (1937) Atomic structure of minerals, 292 p. Cornell University Press, Ithaca, New York. structures of common minerals such as sanidine (Taylor 1933) or ——— (1975) The development of X‑ray analysis. Bell & Sons, London. anorthite (Kempster et al. 1962) represented a significant contri- Bragg, W.H., and Bragg, W.L. (1913) The reflection of X‑rays by crystals. Proceedings of the Royal Society of London A, 88, 428–438. bution to the science of mineralogy. Such breakthroughs continue Bragg, W.L., and Brown, G.B. (1926a) Die struktur des Olivins. Zeitschrift für Kristal- for high-pressure phases; for example the recent demonstration of lographie, 63, 538–566. Al,Si disorder in the octahedral sites of high-pressure Al-rich phase ——— (1926b) Die Kristallstruktur von Chrysoberyll (BeAl2O4). Zeitschrift für Kristallographie, 63, 122–143. D has major implications for water storage and recycling in the Bragg, W.L., and Lipson, H. (1936) The employment of contoured graphs of struc- lower mantle of the Earth (Boffa-Ballaran et al. 2010; Pamato et ture-factor in crystal analysis. Zeitschrift für Kristallographie, 95, 325–337. al. 2015). In addition, the precision and accuracy of modern Bragg Bragg, W.L., and West, J. (1928) A technique for the X‑ray examination of crystal structures with many parameters. Zeitschrift für Kristallographie, 69, 118–148. X‑ray diffraction data can also be used to probe our understanding Brown, G., and West, J. (1927) The structure of monticellite (MgCaSiO4). Zeitschrift of bonding at the fundamental level in stoichiometric compounds, für Kristallographie, 66, 154–161. Brown, G.E., Gibbs, G.V., and Ribbe, P.H. (1969) The nature and variation in length to provide the basic data for explaining the details of minerals at of the Si-O and Al-O bonds in framework silicates. American Mineralogist, 54, the crystal chemical level, and for providing the explanations in 1044–1061. terms of structures of the physical and thermodynamic properties of Brown, P.J., Fox, A.G., Maslen, E.N., O’Keefe, M.A., and Willis, B.T.M. (2006) Intensity of diffracted intensities. In E. Prince, Ed., International Tables for Crystallography, minerals. If used with care, it can be used to provide site occupan- vol. C, p. 554–595. Kluwer, Dordrecht, Netherlands. cies to a precision that starts to become useful for thermodynamic Buerger, M.J. (1931) The crystal structure of marcasite. American Mineralogist, 16, calculations. As a picture of the average structure it provides, in 361–395. Burnham, C.W. (1966) Computation of absorption corrections and the significance of combination with local probes, information about local patterns of end effects. American Mineralogist, 51, 159–167. ordering. This basic idea that the Bragg X‑ray intensities (Eq. 2) Busing, W.R., and Levy, H.A. (1957) High-speed computation of the absorption correction for single crystal diffraction measurements. Acta Crystallographica, represent the time-and-space average of the atomic structure of the 10, 180–182. mineral is also fundamental to studying all other variants on the Busing, W.R., Martin, K.O., and Levy, H.A. (1962) ORNL-TM-305, OR FLS, a Fortran

Downloaded from http://pubs.geoscienceworld.org/msa/ammin/article-pdf/101/5/1036/3603139/5_5473AngelCO.pdf by guest on 27 September 2021 ANGEL AND NESTOLA: A CENTURY OF MINERAL STRUCTURES 1045

crystallographic least-squares program, 83 p. Oak Ridge National Laboratory, in Mineralogy, Mineralogical Society of America, Chantilly, Virginia. Tennessee. Kroll, H., and Ribbe, P.H. (1983) Lattice parameters, composition, and Al/Si order in Capelli, S.C., Burgi, H.-B., Dittrich, B., Grabowsky, S., and Jayatilaka, D. (2014) alkali feldspars. In P.H. Ribbe, Ed., Feldspar Mineralogy, 2, p. 57–100. Reviews in Hirshfeld atom refinement. IUCrJ, 1, 361–379. Mineralogy, Mineralogical Society of America, Chantilly, Virginia. Carpenter, M.A., McConnell, J.D.C., and Navrotsky, A. (1985) Enthalpies of ordering Kroll, H., Lueder, T., Schlenz, H., Kirfel, A., and Vad, T. (1997) The Fe2+,Mg distribu- in the plagioclase feldspar . Geochimica et Cosmochimica Acta, tion in orthopyroxenes: a critical assessment of its potential as a geospeedometer. 49, 947–966. European Journal of Mineralogy, 9, 705–733. Chao, S.H., Hargreaves, A., and Taylor, W.H. (1940) The structure of orthoclase. Min- Kubbinga, H. (2012) Crystallography from Haüy to Laue: controversies on the molecular eralogical Magazine, 25, 498–512. and atomistic nature of solids. Zeitschrift für Kristallographie, 227, 1–26. Cochran, W. (1948) The Fourier method of structure analysis. Acta Crystallographica, Lavina, B., Salviulo, G., and Della Giusta, A. (2002) Cation distribution and structure 1, 138–142. modeling of solid solutions. Physics and Chemistry of Minerals, 29, 10–18. Cole, W.F., Sörum, H., and Kennard, O. (1949) The crystal structures of orthoclase and Matchatschki, F. (1928) Zur Frage der Struktur und Konstitution der Feldspate. Cen- sanidinized orthoclase. Acta Crystallographica, 2, 280–287. tralblatt für Mineralogie Geologie und Palaeontologie Abteilung A: Mineralogie Cox, E.G., and Jeffery, G.A. (1949) The use of ‘Hollerith’ computing equipment in und Petrographie, 97–104. crystal-structure analysis. Acta Crystallographica, 2, 341–343. Meneghinello, E., Alberti, A., and Cruciani, G. (1999) Order-disorder process in the Cruickshank, D.W.J. (1949a) The accuracy of electron-density maps in X‑ray analysis tetrahedral sites of albite. American Mineralogist, 84, 1144–1151. with special references to dibenzyl. Acta Crystallographica, 2, 65–82. Merli, M., Oberti, R., Caucia, F., and Ungaretti, L. (2000) Determination of site pop- ——— (1949b) The accuracy of atomic co-ordinates derived by least-squares or Fourier ulation in olivine: warnings on X‑ray data treatment and refinement. American methods. Acta Crystallographica, 2, 154–157. Mineralogist, 85, 55–65. Dal Negro, A., De Pieri, R., Quareni, S., and Taylor, W.H. (1978) The crystal structures Millikan, R.A. (1910) The isolation of an ion, a precision measurement of its charge, of nine K feldspars from the Adamello massif (Northern Italy). Acta Crystallo- and the correction of Stoke’s law. Science, 32, 436–448. graphica, B, 34, 2699–2707. Nestola, F., Nimis, P., Ziberna, L., Longo, M., Marzoli, A., Harris, J.W., Manghnani, Darwin, C.G. (1914) The theory of X‑ray reflexion. Part II. Philosophical Magazine, M.H., and Fedortchouk, Y. (2011) First crystal-structure determination of olivine 27, 675–690. in diamond: composition and implications for provenance in the Earth’s mantle. Debye, P. (1914) Interferenz von Roentgenstrahlen und Waermebewegung. Annalen Earth and Planetary Science Letters, 305, 249–255. der Physik (Leipzig), 43, 49–95. Onorato, E., Penta, M., and Sgarlata, F. (1963) Struttura del sanidino. Periodico di Della Giusta, A., Carbonin, S., and Ottonello, G. (1996) Temperature-dependent dis- Mineralogia, 32, 1–34. order in a natural Mg-Al-Fe2+-Fe3+-spinel. Mineralogical Magazine, 60, 603–616. Overgaard, J., Schmokel, M., Bjerg, L., Cenedese, S., Jorgensen, M., Chen, Y.-S., and Duane, W. (1925) The calculation of the X‑ray diffracting power at points in a crystal. Iversen, B. (2014) Synchrotron charge density studies of chemical bonding in the Proceedings of the National Academy of Sciences of the USA, 11, 489–493. polymorphs of FeS2. Acta Crystallographica, A, 70, C1338. Farrugia, L. (2014) Accurate H-atom parameters from X‑ray diffraction data. IUCrJ, Pamato, M.G., Myhill, R., Boffa-Ballaran, T., Frost, D.J., Heidelbach, F., and Miyajima, 1, 265–266. N. (2015) Lower-mantle water reservoir implied by the extreme stability of a hydrous Ferguson, R.B., Traill, R.J., and Taylor, W.H. (1958) The crystal structure of low-tem- aluminosilicate. Nature Geoscience, 8, 75–79. perature and high-temperature albites. Acta Crystallographica, 11, 331–348. Pauling, L. (1929) The principles determining the structure of complex ionic crystals. Finger, L., and Prince, E. (1974) A system of Fortran-IV computer programs for crystal Journal of the American Chemical Society, 51, 1010–1026. structure computations. NBS Technical Note 854. National Bureau of Standards, Phillips, M.W., and Ribbe, P.H. (1973) The variation of tetrahedral bond lengths in sodic Gaithersburg, Maryland. plagioclase feldspars. Contributions to Mineralogy and Petrology, 39, 327–339. Finkelstein, G.J., Dera, P.K., Jahn, S., Oganov, A.R., Holl, C.M., Meng, Y., and Duff, Phillips, M.W., Ribbe, P.H., and Gibbs, G. (1973) Tetrahedral bond length variations in T.S. (2014) Phase transitions and equation of state of forsterite to 90 GPa from anorthite. American Mineralogist, 58, 495–499. single-crystal X‑ray diffraction and molecular modeling. American Mineralogist, Ribbe, P.H. (1994) The crystal structures of the aluminum-silicate feldspars. In I. Parsons, 99, 35–43. Ed., Feldspars and their Reactions, p. 1–49. Kluwer, Amsterdam, Netherlands. Friedrich, W., Knipping, P., and Laue, M. (1912) Interferenz-Erscheinungen bei Rönt- Rossi, G., Smith, D.C., Ungaretti, L., and Domeneghetti, M.C. (1983) Crystal chemis- genstrahlen. Sitzungsberichte der mathematiches-physisches Klasse der Königlich try and cation ordering in the system diopside-jadeite: a detailed study by crystal Bayerische Akademie der Wissenschaften zu München., 303–312. structure refinement. Contributions to Mineralogy and Petrology, 83, 247–258. Glazyrina, K., Boffa-Ballaran, T., Frost, D.J., McCammon, C., Kantor, A., Merlini, Sasaki, S., Fujino, K., Takeuchi, Y., and Sadanaga, R. (1980) On the estimation of atomic M., Hanfland, M., and Dubrovinsky, L. (2014) Magnesium and charges by the X‑ray method for some oxides and silicates. Acta Crystallographica, effect of iron oxidation state on its bulk sound velocity at the conditions of the lower A, 36, 904–915. mantle. Earth and Planetary Science Letters, 393, 182–186. Sheldrick, G.M. (2008) A short history of SHELX. Acta Crystallographica, A, 64, Goldschmidt, V.M. (1926) Die Gesetze der Krystallochemie. Naturwissenschaften, 112–122. 14, 477–485. Smith, J.V. (1980) Presentation of the Roebling Medal of the Mineralogical Society of Gruner, J.W. (1929) Structures of sulphides and sulfosalts. American Mineralogist, America for 1979 to W.H. Taylor. American Mineralogist, 65, 578–579. 14, 470–481. Sörum, H. (1953) The structures of the plagioclase feldspars. II. Acta Crystallographica, ——— (1931) Structures of some silicates. American Mineralogist, 16, 437–454. 6, 413–417. Hawthorne, F.C., Ungaretti, L., and Oberti, R. (1995) Site populations in minerals: Stachel, T., and Harris, J.W. (2008) The origin of cratonic diamonds—constraints from terminology and presentation of results of crystal-structure refinement. Canadian mineral inclusions. Ore Geology Reviews, 34, 5–32. Mineralogist, 33, 907–911. Stewart, J.M., Kruger, G.J., Ammon, H.L., Dickinson, C., and Hall, S.R. (1972) The Heinemann, R., Staack, V., Fischer, A., Kroll, H., Vad, T., and Kirfel, A. (1999) X-RAY System of Crystallographic Programs for any Computer, 283 p. University Temperature dependence of Fe,Mg partitioning in Acapulco olivine. American of Maryland, College Park. Mineralogist, 84, 1400–1405. Stewart, R.F. (1976) Electron population analysis with rigid pseudoatoms. Acta Crys- Hejny, C., and Minkov, V.S. (2015) High-pressure crystallography of periodic and tallographica, A, 32, 565–574. aperiodic crystals. IUCrJ, 2, 218–229. Stimpfl, M., Ganguly, J., and Molin, G. (1999) Fe2+-Mg order-disorder in orthopyroxene: Hodgson, M.L., Clews, C.J.B., and Cochran, W. (1949) A punched-card modification equilibrium fractionation between the octahedral sites and thermodynamic analysis. of the Beevers-Lipson method of Fourier synthesis. Acta Crystallographica, 2, Contributions to Mineralogy and Petrology, 136, 297–309. 113–116. Taylor, W.H. (1933) The structure of sanidine and other felspars. Zeitschrift für Kri- Hughes, E.W. (1941) The crystal structure of melamine. Journal of the American stallographie, 85, 425–442. Chemical Society, 63, 1737–1752. Taylor, W.H., Darbyshire, J.A., and Strunz, H. (1934) An X‑ray investigation of the Jones, J.B., and Taylor, W.H. (1961) The structure of orthoclase. Acta Crystallographica, felspars. Zeitschrift für Kristallographie, 87, 464–481. 14, 443–456. Wainwright, J.E.N., and Starkey, J. (1967) Crystal structure of a metamorphic albite. Kempster, C.J.E., Megaw, H.D., and Radoslavich, E.W. (1962) The structure of anorthite, Geological Society of America, Abstracts, 310. CaAl2Si2O8. I. Structure analysis. Acta Crystallographica, 15, 1005–1017. ——— (1971) The refinement of the structure of anorthite. Zeitschrift für Kristallog- Kieffer, S.W., and Navrotsky, A. (1985) Microscopic to Macroscopic: Atomic Environ- raphie, 133, 75–84. ments to Mineral Thermodynamics. Reviews in Mineralogy and Geochemistry, vol Waller, I., and Hartree, D.R. (1929) On the intensity of total scattering of X‑rays. Pro- 14. Mineralogical Society of America, Chantilly, Virginia. ceedings of the Royal Society of London A, 124, 119–142. Kirfel, A. (1996) Cation distributions in olivines and orthopyroxenes. An interlaboratory Wherry, E.T. (1927) The laws of chemical crystallography—review. American Min- study. Physics and Chemistry of Minerals, 19, 503–519. eralogist, 12, 28–31. 2– Kirfel, A., and Will, G. (1980) Bonding in [S2O6] : Refinement and pictorial representa- tion from an X‑ray diffraction study of Na2S2O6·2H2O6 and Na2S2O6·2D2O. Acta Crystallographica B, 36, 512–523. Manuscript received June 22, 2015 Kroll, H. (1983) Lattice parameters and determinative methods for plagioclase and Manuscript accepted August 12, 2015 alkali feldspars. In P.H. Ribbe, Ed., Feldspar Mineralogy, 2, p. 101–120. Reviews Manuscript handled by Martin Kunz

Downloaded from http://pubs.geoscienceworld.org/msa/ammin/article-pdf/101/5/1036/3603139/5_5473AngelCO.pdf by guest on 27 September 2021