<<

Fermionic Superfluidity: From to Stars

Annette Lopez,1, 2 Patrick Kelly,1 Kaelyn Dauer,1 and Ettore Vitali1, 3 1Department of , California State University Fresno, Fresno, California 93740 2Department of Physics, Brown University, Rhode Island 02912 3Department of Physics, The College of William and Mary, Williamsburg, Virginia 23187 From flow without dissipation of to the formation of vortices when placed within a rotating container, the superfluid state of has proven to be a very interesting physical phenomenon. Here we present the key mechanisms behind superfluidity in fermionic systems and apply our un- derstanding to an exotic system found deep within the —the superfluid found deep within a . A defining trait of a superfluid is the pairing gap, which the cooling curves of neu- tron stars depend on. The extreme conditions surrounding a neutron star prevent us from directly probing the superfluid’s properties, however, we can experimentally realize conditions resembling the interior through the use of cold atoms prepared in a laboratory and simulated on a computer. Experimentalists are becoming increasingly adept at realizing cold atomic systems in the lab that mimic the behavior of neutron stars and superconductors. In their turn, computational physicists are leveraging the power of supercomputers to simulate interacting atomic systems with unprece- dented accuracy. This paper is intended to provide a pedagogical introduction to the underlying concepts and the possibility of using cold atoms as a tool that can help us make significant strides towards understanding exotic physical systems.

I. INTRODUCTION superfluidity, and in particular fermionic superfluidity, is described only in very advanced books and papers, which Moving from the “comfort zone” of classical mechanics makes it very hard for a student to have access to this to the more mysterious realm of quantum physics opens very beautiful chapter in physics. We argue on the other the possibility to explore the behavior of at a hand that although the phenomenon appears, or is ex- deeper and very fascinating level. In general, the transi- pected to appear, in some of the most complicated sys- tion from classical to quantum mechanics corresponds to tems in nature, like quantum , the interior of neu- the investigation of microscopic systems: we know that tron stars and superconductors, the underlying physics the motion of matter at the length scale of nuclei and elec- can be captured by some simple and fundamental “in- trons is governed by the celebrated Schr¨odingerequation. gredients”. Nevertheless, in some important cases, quantum effects The purpose of this is to present the key physi- do not remain hidden in the realm of microscopic sys- cal mechanisms underlying superfluidity in a simple but tems but manifest in macroscopic systems, giving rise to precise way, so that students or teachers will be able to plenty of exciting and puzzling physical behaviors that comfortably fill the gap between basic quantum mechani- challenge our intuition and understanding. In fact, the cal problems, like free particles, harmonic oscillators and subtle interplay between quantum mechanics, quantum simple atoms, and more mysterious and fascinating sys- statistics, and interatomic forces give rise to a huge num- tems, like superfluid deep inside a neutron star. ber of fascinating phenomena. We will also shed light into the connection between two A crucial example of such a phenomenon is certainly apparently disconnected fields of physics, and namely superfluidity, together with the closely related supercon- atomic physics and nuclear astrophysics. The term “uni- ductivity. Superfluidity is a unique state of matter char- versality” frequently emerges in physics, and in this con- acterized by the ability of a system to flow without fric- text it may be the key to “reproduce” in a laboratory −9 tion. This results in some spectacular behaviors strongly on earth, at of the order of 10 K, the challenging our common sense, like the observation of conditions that exist deep within a neutron star, where thin films of climbing the walls of a con- the is of the order of hundreds of millions of tainer seemingly defying gravity. Moreover, the super- Kelvins. In a few words, if we put together a collection of 6 fluid state of matter displays other unique properties, for , for example Li atoms in a laboratory or neu- trons in a star, and we have an attractive force acting arXiv:2012.04149v1 [cond-mat.quant-] 8 Dec 2020 example, the formation of quantum vortices when the system is enclosed in a rotating container. Although his- among them, then we are in the conditions to observe a torically superfluidity was first observed in bosonic sys- Fermi superfluid and the behavior of this superfluid will tems, and namely in 4He samples at temperatures below be universal, largely independent from the properties of the celebrated “lambda-temperature”, T = 2.2 K, we the microscopic constituents. We will discuss the foun- now know that the phenomenon can also be observed in dations and we will discuss the evidence that strongly fermionic systems. Even more interestingly, it appears suggests that such a system indeed exists inside a neu- to play a crucial role in some of the most fascinating tron star and it plays a major role in the behavior of the and elusive systems in the universe, like unconventional star itself. superconductors and neutron stars. To our knowledge, This paper is organized as follows: we start with an in- 2 troduction to Fermi , beginning with a description of non-interacting systems and moving to the “univer- sality” of an ultracold interacting , then apply our discussion to the proposed superfluid deep inside a neutron star. We aim to elaborate on the phenomena of glitches in a neutron star as well as to introduce the su- perfluid pairing gap, which the cooling curves of neutron stars are dependent upon.

II. FERMI GASES

Three “ingredients” play a crucial role in fermionic su-

1 perfluidity: quantum mechanics, Fermi statistics and an x attractive force among the fermions. We will now exam- ine these in some detail and we will discuss the emergence of superfluidity from the interplay among the three. x2 From fundamental quantum mechanics, we know how to begin building the mathematical description of the FIG. 1. Illustration of wave function antisymmetry. motion of N -1/2 fermions (for example or neutrons) of mass m. We ultimately want to find dimension. The wave function of the system, Ψ(x1, x2), the wave function of the system, Ψ(r1σ1,..., rN σN , t), where we omit the spin variables to keep the notation 3 where ri ∈ R is the position variable, σi =↑, ↓ is the simple, yields the probability density for the positions of spin variable and t is time. The wave function satisfies the two particles: Schr¨odinger’sequation: 2 |Ψ(x1, x2)| dx1dx2 (4) ∂ i Ψ = Hˆ Ψ (1) ~∂t In other words, it is the probability to observe one neu- tron between x1 and x1 + dx1 and one neutron between where Hˆ is the hamiltonian operator of the system: x2 and x2 + dx2. A fundamental principle in quantum statistics is that identical particles are indistinguishable: Hˆ = Tˆ + Vˆ it is impossible to design an experiment which allows us N to label the particles. This implies that the probability 2 X Tˆ = − ~ ∇2 density must remain invariant if we swap the variables: 2m i i=1 (2) 2 2 |Ψ(x1, x2)| = |Ψ(x2, x1)| (5) N ˆ X V = v(ˆri − ˆrj) When the particle are fermions, in particular, the above i

temperature dependence of the macroscopic properties of the system, for example pressure and energy, is neg- ligible and the system can be described as if it were at T = 0 K. So, while classical systems immediately start to move as we them up starting from T = 0 K, fermionic systems will just “ignore” the change in tem- perature until the temperature approaches the order of the Fermi temperature. Applying this concept can be very counter-intuitive when looking at “natural” Fermi gases, such as neutron stars, since we typically consider them to be “very hot”. When describing the interior temperatures of neutron stars [2], which are on the or- der of 108 K, we observe them to be well below their 12 Fermi temperatures, TF ' 10 K. This huge value is due to the density of the star: in fact, from (7), we see FIG. 2. Illustration of a one-dimensional dispersion rela- that the Fermi energy, and thus the Fermi temperature, tion, showing also the fermions “sitting” on the available mo- depends on the density of the particles. Therefore, we menta.The green dashed lines are occupied energy levels and can consider the interiors of these stars as “very cold” the red circles highlight the Fermi surface (in this case, two even if their actual temperature is hundreds of millions momenta). of Kelvins! In the lab, we can artificially generate ultracold Fermi gases [3] (also referred to as cold atoms) which are way less dense than a neutron star and have a Fermi temper- −8 ature around TF ' 10 K. It is now possible to prepare these gases at temperatures of the order of T ' 10−9 K, so that even these systems can be considered (somehow more naturally) “very cold”. In the next section, we will discover that the two sys- tems, neutrons in a star and cold atoms, share even more properties, other than being both “very cold”: they are both superfluids and their behavior may be surprisingly similar.

C. Interactions and the origins of superfluidity: FIG. 3. In two dimensions, the dispersion relation takes the “universality” of Fermi gases form of a paraboloid. The red circle represents the Fermi surface, the green represents occupied energy levels, and the black represents unoccupied energy levels. We stress that the Fermi sea as we describe it above cannot be a superfluid: there is no way our collection of fermions, “piled-up” like cars in a full parking lot, can flow without friction, which is the defining property of a superfluid. In order to have a flow without friction, we need to have a mechanism that prevents the energy exchange between the system and the walls of the con- tainer in which the system is flowing. In the case of the Fermi sea, energy exchange is readily available: injecting energy into the Fermi sea simply promotes one or more fermions to some unoccupied momenta above the Fermi surface. So, there is no way a simple Fermi sea may be a superfluid: for this, we need to dig deeper by introducing interaction. The model of a non-interacting Fermi gas is extremely useful, and it helps us shed light into the microscopic and macroscopic behavior of plenty of physical systems. But the picture becomes even more intriguing if we al- low the fermions to interact at low temperatures, in par- FIG. 4. Illustration of a three-dimensional Fermi phere. The ticular when the interaction is attractive. The physi- Fermi surface is the surface of the sphere. cal origin of the attraction depends on the details of the 5 physical systems. For example, we know that the gas of will “find” one spin-down partner, and they will bind to electrons in metals experiences an attraction as a con- form a pair. The crucial point is that these pairs, com- sequence of their interaction with the vibrating lattice posite states of two spin 1/2 fermions, act as bosons, and of the positive charges (the nuclei of the metal atoms). the spectacular phenomenology related to Bose statistics While this attraction is normally negligible with respect enters the game. In particular, Bose-Einstein Condensa- to the Coulomb repulsion among the electrons, at very tion (BEC) may take place, leading a macroscopic num- low temperature it becomes important, and it results in ber of pairs to populate a single quantum state [6]. In the phenomenon of : electric current is a fermionic system each quantum state can accommo- able to flow without resistance; superconductivity is the date at most one fermion. However, in a bosonic system, direct analogue to superfluidity, they are two sides of the a macroscopic number of bosons will condense on the same coin. In cold atoms, the attraction is due to the ef- same quantum state when the temperature is very low. fective interaction among the lowest energy states of the An instability takes place here; as fermions start to pair, atoms which can be manipulated using external magnetic the collection of bosonic may condense, and fields. The strong interactions among the neutrons, like a macroscopic number of them will be described by a those in the interior of a neutron star, is attractive in a single-particle (here more properly single-pair) quantum given range of inter-particle distances. wave function. The detailed form of the interatomic potential from (2), v(ri − rj), will naturally be specific to the physical The system behaves in a coherent way [6]: in very sim- system that we are investigating. In general, though, it ple words, a huge number of pairs are doing exactly the will always have a range r0 such that v(ri − rj) ' 0 if same thing! In more precise terms, we have a macro- |ri − rj| > r0: two particles need to be close enough scopic occupation of a quantum state, which thus man- to feel the attractive force. For example, nuclear forces ifests at the macroscopic level. Particles lose their in- are by far the strongest forces that we are aware of, but dividual identities as their wave functions overlap and (fortunately!) we never experience them in our everyday begin to act in unison. This paves the way for superfluid life since their range is as tiny as the size of an atomic behavior. In fact, in order for an energy exchange to nucleus. In most interesting situations involving Fermi happen between the system and its environment, a pair gases, it happens that the average distance between two needs to be broken first. The energy needed to break a V 1/3 pair of fermions is called the superfluid gap of the system, fermions, which can be defined as d = N , is much larger than the range r0. This implies that the “fine” often denoted as ∆, and it is a crucial property of super- details of the interatomic force can be safely neglected. fluids. The gap limits the amount of energy exchange We say that the systems are “dilute”. The possibility to the system has with its environment, therefore, this new ignore most of the details of the interatomic forces allows minimum energy state is now “protected” and can flow us to conclude that dilute gases are “universal”. This without dissipation. allows us to look for similarities between cold atoms and the fluid of neutrons within a neutron star. In fact, both We stress that, when the superfluid is on earth, ∆ systems can be considered to be cold, as discussed above, can be measured experimentally with spectroscopy ex- and also to be dilute. In fact, it is known that in the inner periments (please see for example [7]) and, from the the- crust of a neutron star the average distance among neu- oretical point of view, its calculation is a big challenge. trons [2] is much bigger than the range of nuclear forces, Recently, exact results have started to appear in com- which is comparable to the size of nuclei, of the order putational studies of cold atomic Fermi systems [8, 9]. of 10−15 m. This means that the very complicated, and This energy gap has deep implications for the properties partly unknown, details of the force between two neu- of the system. One physical quantity that is strongly af- trons can be neglected, and the effect of the interaction fected is the specific heat, which can be shown to contain can be taken into account by focusing on the relative a term of the form exp(−∆/kBT ) [10], and has impor- motion of two neutrons when they are very far from each tant consequences for the thermodynamic properties of other. Scattering theory (please see e.g. [4]) dramatically the system. In the realm of neutron stars, for example, simplifies the description of this “dilute” regime: at low the superfluid gap is expected to affect the cooling curves energy, only a few physical quantities are enough to cap- of the stars [10], which describe how the temperature of ture the effect of the interatomic forces. The properties the star evolves as a function of time after the star is of dilute Fermi gases, from cold atoms to neutron stars, born. Very interestingly, if we are able to find a good are “universal” in the sense that they depend only on the estimation of the pairing gap through the analysis of the product kF a [5]. Here, the Fermi momentum kF tells us cooling curves, then we may learn something about the how dense the system is while a is the scattering length superfluid that exists deep in the star, and thus having [3–5] which is a property of the relative motion of two access to properties of matter in a very exotic environ- particles and tells us how strong the interatomic forces ment. The “universality” of Fermi superfluids that we are. discussed earlier would then allow us to mimic these con- The attractive forces may induce the fermions to orga- ditions in a cold gas, mimicking a system that exists in nize in pairs: in the simplest cases, each spin-up fermion the deep universe. 6

D. The wavefunction of the coherent state visible and ultraviolet light, cooling the interior regions. In denser regions, a lack of thermal pressure creates an In the superfluid state, since a very large number of instability, and a perturbation to the cloud, such as a the particles have the same behavior we do not need to shock wave from a supernova or nearby colliding galax- solve for the wave function of a many particle system. ies, can induce collapse. Under the influence of gravity, We have a condensate and thus it is enough to solve for gas will fall to the center of mass. Gravitational po- the wave function of a single boson. A percentage of the tential energy is converted into heat energy under this particles will be in a higher energy state when the system compression, creating a thermal pressure gradient. If the 7 has a finite temperature, but for the sake of simplicity we core temperature reaches 1.5·10 K, nuclear fusion jump- ignore these particles if the temperature is sufficiently starts and the cloud becomes a “protostar”, beginning low. In order to keep the math as simple as possible, we its journey on the main sequence of stars. Thermonu- will disregard the internal structure of a pair of fermions clear fusion sustains the star as thermal and radiation and will treat the pair just like a structureless boson. pressure balance gravity for millions to billions of years. Therefore, the wave function governing the macroscopic If the mass of the star is big enough, the core will burn motion of the superfluid will have the general form: through a chain of elements: hydrogen, helium, carbon, neon, oxygen, magnesium, silicon, and . As each el- Ψ(r) = R(r)eiΘ(r) (10) ement is exhausted, the core contracts until the ignition temperature for the next step of the elemental chain is where ρ(r) = R(r)2 is the probability density for the reached. When iron, the most stable nucleus, is burnt, position of the bosons, while the Θ(r) controls the a critical (and quite dramatic) stage is reached: there current density, that is, the local velocity of the particles. will not be a new fusion reaction. The thermonuclear en- Note that (10) has no dependence on time. In general, ergy is not able to compete against gravitational collapse, the wave function is a function of time, however, here and the star caves-in on itself. The situation inside the we have assumed the system has already relaxed to equi- core during this critical stage is really unique: the en- librium. This equilibrium function is also known as a ergy is such that electrons will combine with the protons steady state. Once the steady state is reached, the state to form neutrons and neutrinos, in a process called in- will not change, unless the system is somehow perturbed verse β-decay. The neutrons will organize in a Fermi sea, and driven out of equilibrium. The wave function (10) which, as we discussed earlier, resists the compression of entirely governs the motion of a superfluid and in partic- the core, making it “stiff.”. Infalling material will thus ular it allows us to understand the formation of quantum rebound, sending out a shock wave. This wave will stall vortices, which is one of the crucial features of superflu- a few hundred kilometers from the center as its energy is ids and are the main reason why it is widely believed dissipated via neutrino losses and photodisintegration of that neutrons deep in a neutron star are superfluid. Be- nuclei. Free-falling material once supported by the core fore discussing this important topic, we will spend some will collide with this stalled shock front to produce an- time “diving” into a neutron star, where one of the most other shock wave that ejects all but the stiffened core in fascinating superfluids exists. a supernovae explosion. As a residual of the explosion, a protoneutron star is born. A protoneutron star has a temperature on the order of 10 MeV, and escaping neu- III. THE FORMATION OF A NEUTRON STAR trinos will cool the star to an MeV or less during the first thousand years of life. Now that we have presented the key ingredients that allow us to understand the physical mechanisms under- lying the spectacular phenomenology of superfluids, we IV. THE STRUCTURE OF A NEUTRON STAR find it a beneficial example for deepening our understand- AND THE NEUTRON SUPERFLUID ing to discuss one of the most mysterious superfluids that exists in nature: the fluid of neutrons in a neutron star. Glancing at a neutron star, one will immediately no- We will certainly need some context, so we will briefly tice a hydrogen atmosphere. Directly beneath will lie provide a general description of the stars, starting from a thermal , known as the envelope, which acts their birth. Later, we will discuss how the superfluid as a barrier between the hot interior and the surface at- influences, or is speculated to influence, the behavior of mosphere. Diving deeper, we expect to see four internal the star through a description of glitches and the cooling regions of the neutron star that are distinguishable by process. nucleon densities: the outer crust, the inner crust, the Interstellar gas clouds are diffuse and primarily com- outer core, and the inner core (depicted in figure 5). The posed of hydrogen molecules and dust (particles made outer crust is a few hundred meters below the envelope of carbon, silicon, and oxygen atoms). These gaseous and is expected to be a nuclear lattice composed of heavy conglomerates can span distances from less than a light- nuclei. The inner crust contains a lattice of neutron rich year to several hundred light-years. We see the temper- nuclei in equilibrium with a superfluid of neutrons which 1 ature vary inside the interstellar clouds: dust absorbs are paired in the S0 quantum singlet state and a de- 7 generate gas. Here the spectroscopic notation 1 S0 simply means that the neutrons in the inner crust form pairs made of one spin-up and one spin-down, like we have discussed in the previous sections. As we ven- ture into the outer core, the system becomes more and more exotic: we do not have nuclei any more; instead, a we have a fluid of free neutrons, protons, and electrons. b It is assumed that both the neutrons and the protons c are superfluids. The proton fluid is charged and would d actually be considered a Type II superconductor. The e neutrons would pair in a triplet state—two neutrons with the same spin orientation would pair—while the protons 1 would form pairs in the more “conventional” S0 singlet state. A “soup” of two Fermi superfluids of protons and neutrons is thus expected to form the deepest region of the neutron stars: the interaction between the two com- ponents and the extreme density can make the physics very challenging. The inner core is mysterious because its composition and are unknown to the scientific community, however, there are several proposi- tions for the composition of the core matter: nucleons; FIG. 5. Structure of a neutron star: (a) the atmosphere and pion or kaon condensates; hyperons; and . For the envelope are the outermost layer, followed by (b) the outer purpose of this paper, we focus on the superfluid of neu- crust and (c) the inner crust, where the superfluid transition trons in the inner crust, which is “dilute” and universal, occurs, and (d) the outer core and (e) the inner core. in the sense discussed in section II C; if we knew the in- teraction strength, or the superfluid gap, we would be tainer is a cylinder that is rotating around its axis, which able to “observe” it in a laboratory in the form of a cold we chose to be the z axis of our reference frame. It is nat- atomic Fermi gas. Cold atoms would “impersonate” the ural, simply by transforming the energy into a rotating neutrons and we would be able to tune the interaction reference frame, to expect that the wave function ψ(r) strength to reproduce the physics inside the mysterious will be an eigenvector of the component of the angular star. This may be achieved through observation of the momentum along z, Lˆz, implying that Θ(r) = mϕ, where cooling curves of the stars, which would allow us to infer m is an integer number, while ϕ is the cylindrical angular the value of the interaction strength. Before discussing coordinate of the position r = (x, y, z): ϕ = arcsin(y/r), this very important point, we would like to discuss the with r = px2 + y2 being the distance from the axis. The evidence for the existence of a neutron superfluid inside relation (11) leads us to the interesting result: a neutron star: the phenomenon of glitches. 1 v(r) ∝ e (12) r ϕ V. VORTICES AND GLITCHES where eϕ is the unit vector along ϕ. This velocity field describes a vortex, as seen in figure 6. So, the fact that a We now provide a short mathematical discussion of the superfluid is a coherent state of a macroscopic number of superfluid vortices inside the rotating crust of a neutron pairs described by a single-pair quantum wave function star. As long as the rotation is slow, the superfluid will naturally allows us to conclude that, within a rotating remain steady, since the container is not able to drag container (provided that the rotation is fast enough), the the nonviscous fluid. A general expression for the local superfluid will form vortices. From thermodynamic ar- velocity, given the wave function (10), is guments it may be seen that, the higher the speed or rotation, the greater is the number of vortices. ~ v(r) = ∇Θ(r) (11) A detailed description of the origin of neutron star m glitches would certainly require a highly sophisticated which corresponds to the velocity field of the conden- presentation, but we can capture the essence of the sate. We will derive this expression in the appendix. For phenomenon with a few simple ingredients: a fluid of a steady or slowly rotating superfluid, the phase Θ(r) fermions (the neutrons) attract each other, form pairs, does not depend on the position and thus we have no net and as the star is rotating, form vortices. If the star is velocity field. The situation does change when the rota- rotating (rotating neutron stars are typically called pul- tion becomes fast. The wave function that describes the sars), the neutron superfluid inside a neutron star is ex- motion of the system will correspond to the minimum en- pected to host an array of vortices. The fact that the su- ergy when studied from a reference frame that is rotating perfluid coexists with a lattice of nuclei makes it possible with the container. Imagine, for simplicity, that the con- for the vortices to be pinned to the lattice sites, similar to 8

such that the core and crust reach thermal equilibrium. Within the period of 10 − 100 years < t < 105 − 106 years, neutrino emission continues to dominate. Stage three refers to a mature neutron star experiencing pho- ton emission via the transport of heat from the core to its surface. From 105 − 106 years and onward, the evolution of the core temperature is thus governed by this radia- tion. The surface temperature of a star is dependent on the age of the star, and this relationship is shown with cooling curves. Cooling curves depend strongly on the properties of matter within the core of the neutron star, and are heav- ily dependent on the local density in the interior of the star and, in particular, they are sensitive to the existence of a superfluid. For a non-superfluid core, the transition from stage one to stage two consists of a sharp off in temperature due to the direct URCA process. A direct URCA process consists of a pair of reactions:

n → p + e− + ν FIG. 6. Top-view diagram of the vortex velocity vector flow e − (13) direction field in (12). p + e → n + νe vortices in rivers that are pinned to rocks on the riverbed. where n is a neutron, p is a proton, e− is an electron, The rotation of the star gradually slows down as the star and νe is an electron neutrino. The emitted neutrinos emits energy. As this happens, the vortices will feel a are responsible for the cooling of the star, as they lit- force, known as a Magnus force, that pulls them towards erally transport energy away from the star. These re- the surface of the star. This force can break the pinning, actions are possible given the right condition of density, and the vortex may hit the surface. As this happens, a ρ ≥ 4.62ρnuclear, where the nuclear density is ρnuclear = glitch takes place—the star spins faster for a short time— 2.3×1017kg/m3 which is possible for stars whose mass is and observations of these glitches give indirect evidence MNS ≥ 1.35M , where M is the mass of the sun. This of a superfluid core. It is quite amazing to see how the is because the resource of proton concentration must be simple physical mechanisms discussed in this paper al- sufficient (a ratio of the number of protons to the number low us to give a possible explanation to a phenomenon of nucleons needs to be approximately 0.1). that happens in the deep universe. We comment that In contrast, superfluidity suppresses the neutrino lu- other models have been proposed to explain the glitches, minosity, slowing down the cooling in stage two. This for example star-quakes, but to our knowledge the role is due to the fact that nucleons must be excited above of the superfluid vortices and their unpinning is widely the pairing gap (discussed in section II C) inherent in a accepted in the astrophysics community as the leading superfluid to participate in the direct Urca process. A mechanism. superfluid core will cool the star in stage one and stage two such that the observable crust temperature is lower than the critical temperature necessary for superfluidity. VI. COOLING OF NEUTRON STARS Because of this, neutron stars are ideal for determining the critical temperatures of superfluid matter at super- While the glitches are certainly the most spectacular nuclear densities. Interestingly, an accurate analysis of manifestation of the superfluid existing deep inside a neu- the may yield an estimate of the superfluid tron stars, the superfluid neutrons are expected to affect gap ∆, which, in turn, may give us information about the also other properties of the star, and in particular the Fermi superfluid, and namely the interaction strength. cooling process. A detailed description can be found in This information, thanks to the universality of Fermi su- [? ], and we just give a brief and non-exhaustive descrip- perfluids, could allow us to mimic the neutron superfluid tion here, in order to allow the reader to understand the using cold atoms, as we will discuss in the next section. role played by the superfluid. Neutron stars undergo three main cooling stages. In stage one, the crust is thermally decoupled from the inte- VII. EXPERIMENTAL ANALYSIS OF THE rior such that the surface temperature of the star reflects SUPERFLUID PAIRING GAP OF A NEUTRON the thermal state of the crust. In the early life of a neu- STAR tron star, 10 − 100 years, cooling primarily happens via neutrino emission. During stage two, the surface temper- In the last few decades, after the celebrated realization ature of the crust will adjust to the internal temperature, of Bose-Einstein- in 1995, the possibility of 9 cooling down collections of atoms to temperatures very close to the opened the unique possibility to produce Fermi (and Bose) gases in a laboratory. In the study of fermions, typically Lithium or Potassium atoms are cooled down with advanced cooling and trap- ping techniques. The resulting collections of very cold particles are realizations of quantum gases which can be studied with unprecedented experimental control. For example, we can control the interatomic forces by simply tuning an external magnetic field using a phenomenon known as Feshbach resonance. We are literally able to engineer a superfluid with a given value of kF a, as dis- cussed above.

For such “artificial superfluids”, we can access the su- FIG. 7. Color plot of the spectral function of a dilute attrac- perfluid gap ∆ and we can study its dependence on kF a, tive Fermi gas. Source: Annette Lopez and Ettore Vitali that is on the density and on the strength of the interac- tion. This can be done both experimentally and theoret- ically, thanks to the unprecedented progress in compu- tational methodologies that we witnessed in the last few decades. From the experimental point of view, the study of the pairing gap relies on spectroscopy experiments, which measure the so called spectral function A(k, ω), yielding a map of the quantum states that are available for the particles in the system. The same function can be also estimated using theoretical and computational tech- niques, and we show an example in figure 7. The zero of the energy is at the Fermi surface, or more precisely at the chemical potential corresponding to the given parti- cle density. For ω < 0, the spectral function tells us the probability that a particle with a certain momentum k has an energy ω (here, in units of the Fermi momentum and Fermi energy). The broadening is the corresponding FIG. 8. Left: dispersion curve for the non-interacting system. uncertainty, arising from the fact that the momentum of Right: dispersion curve of an interacting system. The inter- one particle is not a good quantum number for an inter- acting system has a prohibited region where no fermions may acting system: a particle with a given momentum k can occupy energy states known as the superfluid pairing gap, ∆ have an energy that is distributed around a mean. For (highlighted in orange is 2∆ symmetric about 0 energy). ω > 0, the spectral function shows us the available states relationship for new particle that we may inject into the system in p 2 2 2 a spectroscopy experiment. From figure 7 we clearly see ω = ± (k − EF ) + ∆ (14) a gap between the “occupied” states at ω < 0 and the “available” states at ω > 0: this is exactly the superfluid where ∆ is the gap. An interacting system has two gap. branches, separated by the energy gap required to break apart a pair of fermions. This gap is the superfluid gap The function ω(k), giving the mean energy as a of the system: the gas can flow without dissipation as function of the momentum and capturing the peaks in long as the the container is unable to break the pairs. A(k, ω), is called the dispersion relation of the system. In With cold atoms, like Lithium or Potassium, we can figure 8 we plot the dispersion relations for both a non- study these superfluids on earth and learn the depen- interacting Fermi gas, and for an attractive Fermi gas. In dence of the superfluid gas, and of other properties of both cases the zero of the energy is set to the Fermi en- the superfluid, like the vortices, on the interatomic forces, ergy (or chemical potential for the interacting system), which can be controlled with high accuracy. Also, we can so that negative ω corresponds to occupied states and sharpen our theoretical and computational approaches to positive ω corresponds to unoccupied states. In the non- give more and more accurate predictions. Very interest- interacting case, no energy is expended for a particle to ingly, these systems are somehow universal, as discussed move into a higher energy state. If we juxtapose the above, and so what we learn in a cold atomic system can non-interacting dispersion curve with an interacting dis- tell us something about the superfluid neutrons deep in- persion curve, shown in figure 8, we can clearly see the side a neutron star. A very exciting research direction presence of a gap, visible around zero energy. In the inter- is as follows: if we can infer the value of the superfluid acting case, the dispersion curves follow the approximate gap from the cooling curves, as discussed above, can we 10 learn more about the interaction strength and, maybe, gap of the superfluid core. Ultracold atomic Fermi sys- “reproduce” the superfluid in a laboratory on earth? tems are a tool we can utilize to compute the pairing gap of a superfluid, which can be compared to models of cooling curves, shedding light on the interior of neutron VIII. SUMMARY AND CONCLUSIONS stars. Science is ultimately about observations and experi- ments on real physical systems. Experiments tend to be Using quantum mechanics we can describe the na- expensive and difficult, and thus the number of experi- ture of Fermi gases. Non-interacting fermions are fully ments and observations a researcher would like to do is governed by the Pauli exclusion principle which deter- much smaller than the number of experiments they ac- mines the organization of the Fermi sea: fermions will tually can do. In addition, for practical reasons many occupy all energy states up to the density dependent systems in nature are impossible to study directly, such Fermi energy. In the quantum realm, each particle is as most astronomical phenomena. Obviously the era of in a plane wave, with a given momentum. Looking at space exploration has allowed for us to have direct con- a non-interacting Fermi gas in momentum space, we see tact with the greater universe, but for the foreseeable that if any amount of energy is introduced to the sys- future we are going to be confined within the solar sys- tem, fermions can easily use it to occupy states above tem. The nearest neutron star is approximately 400 light the Fermi energy. Fermi gases have internal tempera- years away, roughly 25 million times the distance between ture scales dependent upon the Fermi temperature (di- the Earth and the Sun. The technical challenge of recre- rectly related to the Fermi energy). So long as the Fermi ating certain systems in the laboratory is also a major gas’s temperature is lower than the Fermi temperature, impediment to experimental science. Rapid progress is the system can be described as if held at absolute zero. being made in realizing superfluid and phases As quantum effects become negligible when the temper- in cold atomic gas systems. Because of the universality ature reaches the Fermi temperature, we can say that described above, these may serve as useful real models of all quantum Fermi gases are “cold”. If we allow the the superfluid interiors of neutron stars. fermions to interact, if the average distance between the We want to stress the importance of computational the fermions is much larger than the range of the force physics for investigating systems such as quantum fluids. resulting from the interparticle potential, we can also de- Purely theoretical investigations are extremely difficult clare a Fermi gas as being “dilute”. Cold and dilute Fermi for complex systems. Computational physics can be seen gases are “universal”. This ”universality” ultimately al- as an implementation of theoretical physics at a much lows us to compare and connect two systems literally larger scale than can be done with pencil and paper work lightyears apart from each other: cold atoms made in a alone. Furthermore many of the methods implemented laboratory and the superfluid cores of neutron stars. by computational physicists were invented before mod- An attractive interacting potential amongst the ern computers existed or were powerful enough to tackle fermions will result in pairs of fermions, which ultimately physically meaningful systems. Computational physics behave as bosons. These integer spin pairs no longer obey will not be a replacement for experimental or theoretical the Pauli exclusion principle and ultimately all collapse work. However it combines the inexpensiveness of theory into the ground state, leading to a macroscopic mani- with the ability to account for the greater complexity of festation of a quantum phenomenon—superfluidity—in real physical systems. Computational results also serve which the pairs of fermions behave “coherently”. In sim- as useful benchmarks that provide a roadmap for more ple words, they all do the same thing! In more rigorous ambitious experiments within the next decade. terms, a macroscopic number of fermions is described by the same single-pair “collective” wave function. A char- acteristic trait of a superfluid is the pairing gap: the Appendix A: Derivation and Motivation for the energy required to break a pair of fermions. The energy Probability Current states above the fermi energy become prohibited unless energy greater than or equal to twice the pairing gap is Here we would like to provide a motivation for (11). injected into the system. This is what allows a superfluid The explanation is straight forward but without it, this to flow without friction. Another characteristic trait is equation can seem to have appeared from nowhere. At the formation of quantum vortices, when the system is first sight, the exclusive dependence on the phase can be in a rotating vessel, like the crust of a spinning neutron quite perplexing. To understand this equation will start star. by deriving an expression for the probability current, and Observations of glitches provides indirect evidence of from there find (11). Suppose we have a region of space the existence of a superfluid in the deep universe: a rapid V and a particle with wave function Ψ(r, t). We would acceleration in the rotation rate of a neutron star. These like to know how the probability of finding the particle are predicted to be a result of the quantized vortices in V changes with time. We write the probability as which can form within a rotating superfluid. Theorists have also concluded that the cooling of neutron stars, Z P (V) = ρ d3r (A1) from birth to death, have a dependence on the pairing V 11

Where the probability density is ρ = R2 = |Ψ|2 = Ψ∗Ψ. which intuitively means that if a fluid moves from one Taking the time derivative of the density we have region to another, it must move through all the regions in-between. This means any concentration of probability ∂ρ ∂Ψ∗ ∂Ψ = Ψ + Ψ∗ (A2) in one region will be at the expense of probability flowing ∂t ∂t ∂t out of adjacent regions. Substituting the Schr¨odingerequation (where we assume We now want to obtain (11). We start by writing the we are working in units where ~ = 1) wave function in polar form using (10) and substituting this into the probability current given in (A5). This gives ∂Ψ 1 i = − ∇2Ψ + V Ψ (A3) ∂t 2m i J = ReiΘ∇Re−iΘ − Re−iΘ∇ReiΘ (A8) (and its complex conjugate) into (A2) we obtain 2m ∂ρ i = − (Ψ∇2Ψ∗ − Ψ∗∇2Ψ) = −∇ · J (A4) ∂t 2m Where we have suppressed the ~q and t dependence. Eval- uating the gradients and simplifying we obtain Here we have defined a new quantity, the probability cur- rent ρ i J = ∇Θ(r, t) (A9) J = (Ψ∇Ψ∗ − Ψ∗∇Ψ) (A5) m 2m

Using (A1) and (A4) and applying the divergence theo- In classical hydrodynamics the current is given as J = ρv. rem, we can write Here ρ is the density of a fluid parcel moving with velocity Z Z v. We can rewrite this expression in terms of momentum ∇ · Jd3r = J · nd3r (A6) p as V ∂V Where ∂V is the surface surrounding the region V. ρ J = p (A10) Intuitively, this means we can treat the probability m density like an inhomogeneous fluid, that is, a fluid with Equating (A9) and (A10) we have a variable density, like a gas. As time passes in some re- gions of space probability can become more rarefied and 1 in other regions probability can accumulate. The proba- v(r, t) = ∇Θ(r, t) (A11) bility is a conserved quantity (the total amount is always m equal to one) and obeys the continuity equation

∂ρ Which gives us (11): the velocity field of a particle in the = −∇ · J (A7) superfluid. ∂t

[1] R. Pathria and P. D. Beale, in [6] R. Pathria and P. D. Beale, in Statistical Mechanics (Third Edition), edited by Statistical Mechanics (Third Edition), edited by R. Pathria and P. D. Beale (Academic Press, Boston, R. Pathria and P. D. Beale (Academic Press, Boston, 2011) third edition ed., pp. 231 – 273. 2011) third edition ed., pp. 179 – 229. [2] S. L. Shapiro and S. A. Teukolsky, [7] F. Schwabl, Advanced Quantum Mechanics (Springer, Black Holes, White Dwarfs, and Neutron Stars (John Berlin, Heidelberg, 2004). Wiley and Sons, Ltd, 1983) Chap. 10, pp. 267–305. [8] E. Vitali, H. Shi, M. Qin, and S. Zhang, Phys. Rev. A [3] S. Giorgini, L. P. Pitaevskii, and S. Stringari, Rev. Mod. 96, 061601 (2017). Phys. 80, 1215 (2008). [9] E. Vitali, P. Kelly, A. Lopez, G. Bertaina, and D. E. [4] D. J. Griffiths and D. F. Schroeter, Galli, Phys. Rev. A 102, 053324 (2020). Introduction to Quantum Mechanics, 3rd ed. (Cam- [10] S. L. Shapiro and S. A. Teukolsky, bridge University Press, 2018). Black Holes, White Dwarfs, and Neutron Stars (John [5] A. Gezerlis and J. Carlson, Phys. Rev. C 77, 032801 Wiley and Sons, Ltd, 1983) Chap. 10, pp. 267–305. (2008).