Brought to you by:
Review

Kosterlitz–Thouless physics: a review of key issues

Published 28 January 2016 © 2016 IOP Publishing Ltd
, , Citation J Michael Kosterlitz 2016 Rep. Prog. Phys. 79 026001 DOI 10.1088/0034-4885/79/2/026001

0034-4885/79/2/026001

Abstract

This article reviews, from a very personal point of view, the origins and the early work on transitions driven by topological defects such as vortices in the two dimensional planar rotor model and in 4Helium films and dislocations and disclinations in 2D crystals. I cover the early papers with David Thouless and describe the important insights but also the errors and oversights since corrected by other workers. I then describe some of the experimental verifications of the theory and some numerical simulations. Finally applications to superconducting arrays of Josephson junctions and to recent cold atom experiments are described.

Export citation and abstract BibTeX RIS

1. Introduction

From my point of view, the discovery of defect mediated phase transitions started in 1970 when I was a postdoc in high energy physics at the Istituto di Fisica Teorica of Torino University in Italy. I was, and still am, a disorganized person with a tendency to wait until the last possible moment or even later to do something important like submitting a job application. Consequently, in September 1970, I found myself in the Department of Mathematical Physics, Birmingham University, England instead of at CERN in Geneva, Switzerland where I had intended to go. During my first year at Birmingham, I did a set of complicated calculations on a model which was a precursor to string theory, but I was always beaten by a group in USA. At least twice, I was in the process of writing up my calculations when a preprint by my competitors appeared in the department.

Somewhat discouraged, I began to look at other branches of physics in the hope of finding an interesting but tractable problem and found myself talking with David Thouless about strange excitations such as vortices in superfluid 4Helium films and domain walls in the Ising model of magnetism. This opened my mind to the world of condensed matter physics where, to my uneducated mind, there seemed to be a plethora of unsolved physical problems. This was a big contrast to high energy physics where there seemed to be very few interesting problems with several very smart people working on every problem. However, there was another difficulty which was my almost total ignorance of statistical mechanics which I had more or less ignored as a graduate student in high energy physics.

I had been warned by my colleagues that David could be difficult because he was reputed to not suffer fools gladly. There was little question that I was an ignorant fool as far as phase transitions were concerned so I was quite nervous listening to David's thoughts about some low dimensional problems in phase transitions. He explained that he had been discussing with Phil Anderson the problem of a phase transitions in the 1D Ising chain with long range interactions between spins falling off as 1/r2. He had recently shown [1] that, in this model, there was a phase transition with some peculiar properties. It was known that an Ising chain with shorter ranged interactions had no transition to an ordered state at any finite temperature [2] while the system with interactions decaying more slowly than 1/r2 had an ordered state at sufficiently low temperatures [3] and, thus, a transition at a finite T  >  0.

Our early discussions of this and related problems introduced me to more new concepts like topological defects such as vortices in superfluid 4He and their role in disordering a system. As I was operating from a position of extreme ignorance, these did not seem any stranger than other ideas about phase transitions. In fact, they seemed more hopeful than many others all of which had failed to make much of an impression on this class of intractable problems. In an Ising model, a topological defect is an easily visualized excitation as it is just a domain wall between regions of opposite spin orientation. It is not difficult to rewrite the Ising spin Hamilton in terms of interacting domain walls which live on the lattice dual to the original spin lattice. During these discussions, David gave me a set of papers by Anderson and coworkers [48] and suggested I look at them as they may be useful. These were a major surprise because, as far as I was concerned, their importance was that they solved the 1D Ising model with 1/r2 interactions and demonstrated that it is ordered at low temperatures and has a phase transition to a disordered state at a finite ${{T}_{\text{c}}}$ . I spent most of six months doing nothing but reading and rereading these papers until I understood what they were doing, especially the last in the series [6] which solved the problem by a very strange method. I later realized that this paper contained a very early version of the renormalization group popularized by Wilson [9, 10] to formalize the scaling theory of Kadanoff [11] but, at the time, it just appeared to be a very clever and unusual method of solving one particular example of a phase transition.

1.1. One dimensional Ising model

In 1969, Thouless was first exposed to the essential ideas vital to theories of transitions driven by topological defects during a visit to Bell Labs, one of the most important centers of theoretical physics in the twentieth century. There he learned of the work by Anderson and colleagues on the transformation of the Kondo problem into a 1D Ising model with 1/r2 interactions [48] and was asked if he knew anything about this intermediate case. It was known that the model with interactions falling off more slowly than this have long range order [3], or finite magnetization, at low temperatures and that systems with shorter ranged interactions are disordered [2] at all T  >  0.

To answer this challenge, Thouless constructed an argument [1] based on that of Peierls [12, 13] and Landau and Lifshitz [14]. In it, one considers a system of length L with a single domain wall between spins with s  =  +1 and spins with s  =  −1. The entropy is just

Equation (1)

and the energy of this isolated domain wall at x is

Equation (2)

for the domain wall at $0\ll x\ll L$ well away from the ends of the chain. Thus, the free energy $\Delta F$ of an isolated domain wall is

Equation (3)

The probability of a single domain wall in a system of length $L\gg a$ is

Equation (4)

This implies that one can identify the critical temperature as ${{k}_{\text{B}}}{{T}_{\text{c}}}=J$ since there will be no isolated domain walls in the equilibrium system when $T<{{T}_{\text{c}}}$ so that the system will be ordered. There will be finite domains of reversed spins below ${{T}_{\text{c}}}$ but still a finite magnetization. On the other hand, for $T>{{T}_{\text{c}}}$ , there is a finite probability that there will be some isolated free domain walls implying disorder and vanishing magnetization. The original argument [1] went further to show that the magnetization has a discontinuity at ${{T}_{\text{c}}}$ although the transition is continuous. This was later confirmed by more rigorous work [1517].

The free energy $\Delta F$ of equation (3) for an isolated domain wall in the 1/r2 ferromagnetic Ising chain has important implications which were to become relevant to later developments in the defect theory of phase transitions. Domain walls can be regarded as topological defects in an Ising ferromagnet and it is obvious that a finite concentration of these in thermal equilibrium tend to disorder the system. Equation (4) shows that there will be no isolated domain walls in the system for ${{k}_{\text{B}}}T<J$ and there will be long range ferromagnetic order. On the other hand, when ${{k}_{\text{B}}}T>J$ , a finite density of these topological defects is present in thermal equilibrium thus disordering the system. This argument is made more exact by Anderson, Yuval and Hamann [6, 8] who treated the 1/r2 Ising chain by a renormalization group method. Because this very early version of a renormalization group treatment of a defect driven phase transition was so important to later developments by Thouless and myself [1820], I review it here.

The first step is to rewrite the partition function Z(K, y, h) of the 1D Ising model with Hamiltonian [6, 8], assuming periodic boundary conditions

Equation (5)

in terms of a set of 2n alternating domain walls at ${{r}_{1}},\cdots {{r}_{2n}}$ where $\Delta s=\pm 2$ across a wall. The partition function becomes [6, 8]

Equation (6)

where ${{q}_{i}}={{(-)}^{i}}$ , ${{K}_{0}}(T)=\frac{J}{{{k}_{\text{B}}}T}$ and ${{y}_{0}}=\text{exp}\left(-\frac{A}{{{k}_{\text{B}}}T}-{{K}_{0}}(T)(1+\gamma )\right)$ , with $\gamma =$ Euler's constant, is the domain wall fugacity, which is a measure of the defect density. As will become apparent, the value of the fugacity y0 affects only the value of the transition temperature ${{T}_{\text{c}}}$ but not its nature. Note that A, the nearest neighbor coupling constant, affects only the the domain wall fugacity y and not the nature of the transition, provided A  >  0. The idea is to proceed perturbatively [46, 8] and an expansion parameter is needed. A system of very dilute domain walls is amenable to an expansion in powers of the fugacity y0 which can be made small by choosing $A\gg 1$ and this expansion is the basis of the renormalization group procedure. This is a somewhat unusual expansion as ${{y}_{0}}\sim {{\text{e}}^{-\mu /{{k}_{\text{B}}}T}}$ is very singular at T  =  0 so, although this expansion is most reasonable at low T, it is not a standard low temperature expansion.

The partition function $Z\left({{K}_{0}},{{y}_{0}},h\right)$ defined by (6) is not analytically accessible but information can be obtained by a renormalization group procedure [5, 6, 8]. This consists of doing a partial trace by integrating over the domain wall positions in the interval $a<{{r}_{i+1}}-{{r}_{i}}<a{{\text{e}}^{\delta l}}$ and rescaling the short distance cutoff or lattice spacing a to obtain a new partition function describing fluctuations on length scales larger than $a(l)=a{{\text{e}}^{l}}$ . The rescaled Hamiltonian has the same form as the original Hamiltonian with cutoff a(l) provided the interaction parameters ${{K}_{0}},{{y}_{0}},h$ are replaced by renormalized parameters K(l), y(l), h(l) which are solutions of the renormalization group (RG) flow equations for K(l), y(l), h(l) and the free energy per unit length f(l), [6, 8, 21]

Equation (7)

It was a major revelation in 1971 when I understood that, by integrating these flow equations and by using a bit of physical intuition, it was possible to extract the behavior of some physically relevant quantities such as the magnetization, susceptibility, specific heat and the correlation length. This was fortunate as calculating the partition function $Z\left({{K}_{0}},{{y}_{0}},h\right)$ was well beyond my ability, but Anderson's scaling ideas provided a more feasible method to obtain information. The flow equations of (7) can be derived with some difficulty but within a finite time and integrating these gives

Equation (8)

where the integration constant C is also a renormalization group invariant.

The transition temperature ${{T}_{\text{c}}}$ is determined by equation (8) with C  =  0 which gives

Equation (9)

which is the dashed line in figure 1. The solid line which is essentially indistinguishable from the dashed line is ${{T}_{\text{c}}}$ from (8) with C  =  0.

Figure 1.

Figure 1. The critical temperature ${{T}_{\text{c}}}$ as a function of y0. The solid line from (8) with C  =  0 and the dashed line from (9) are indistinguishable for y0  <  0.3.

Standard image High-resolution image

The flows are plotted in figure 2 and the arrows denote the direction of increasing l. $C\left({{K}_{0}},{{y}_{0}}\right)$ is a constant of integration which depends on ${{K}_{0}}=J/\left({{k}_{\text{B}}}T\right)$ and the fugacity y0, defined below equation (6). Since the flow line separating regions I and II flows into the point $K(\infty )=1,\,y(\infty )=0$ , all systems with initial parameters ${{K}_{0}},{{y}_{0}}$ lying on this separatrix can be interpreted as being at the critical temperature ${{k}_{\text{B}}}{{T}_{\text{c}}}\left(\,{{y}_{0}}\right)/J=K_{\text{c}}^{-1}\left(\,{{y}_{0}}\right)$ . For $T={{T}_{\text{c}}}(1+t)$ , C  =  −t(4y0  −  t) to lowest order in t and y0. Systems in region I flow to some point on the fixed line at $K(\infty )\geqslant 1,\,y(\infty )=0$ which corresponds to a completely ordered system, while a system in region II or III flows to one where y(l) increases to $\mathcal{O}(1)$ . This equivalent system has a large density of domain walls which is interpreted as complete disorder or $T>{{T}_{\text{c}}}$ . The RG flows are sketched in figure 2 and the transition temperature as a function of y0 in figure 1 from (8) with C  =  0.

Figure 2.

Figure 2. Flows of the RG equations, (7) in regions I, II, III. The flows with increasing l are in the direction of the arrows.

Standard image High-resolution image

One of the big surprises was that the zero field magnetization m has a finite value as $T\to T_{\text{c}}^{-}$ so that there is a discontinuity in $m(T)=m\left({{T}_{\text{c}}}\right)\left(1-b\sqrt{{{T}_{\text{c}}}-T}\right)$ where b is a positive constant [8]. It is interesting to note that this discontinuity in the spontaneous magnetization was first predicted by Thouless [1] and derived from the RG flow equations (7) [8]. It was finally proved rigorously [16] that the model has a finite spontaneous magnetization at low T and later that the magnetization m was discontinuous at the transition [17].

Since the form of the RG flow equations (7) turned out to be very similar to those of the 2D planar rotor model [20], I rederive some of the important results of Anderson and Yuval [8] in more modern language. One of the most remarkable features of the Ising ferromagnet of (5) is that the magnetization jumps discontinuously to zero at ${{T}_{\text{c}}}$ . This follows from the solution to the RG equations. Linearize (7) about K  =  1 by writing K  =  1  +  x so that to $\mathcal{O}(x)$ ,

Equation (10)

Here $C(t)=x_{0}^{2}-4y_{0}^{2}=\left({{x}_{0}}-2{{y}_{0}}\right)\left({{x}_{0}}+2{{y}_{0}}\right)$ , the small $\left({{x}_{0}},{{y}_{0}}\right)$ expansion of C of equation (8), is a constant of integration which depends on the deviation t from the critical temperature ${{T}_{\text{c}}}$ . The RG flows are a set of hyperbolae in the x, y plane in figure 3. The transition is at x  =  y  =  0 and the flows are towards the line $x(\infty )\geqslant 0,y(\infty )=0$ when $C(t)\geqslant 0$ and ${{x}_{0}}\geqslant 2{{y}_{0}}$ , which defines region I. When C(t)  <  0, y(l) becomes relevant and flows to $y(l)=\mathcal{O}(1)$ , defining region II. Finally, in region III where $C(t)\geqslant 0$ and ${{x}_{0}}\leqslant -2{{y}_{0}}$ , the fixed line y  =  0 is unstable and y(l) is relevant, increasing with l. In particular, one identifies the transition as C  =  0 when x(l)  =  2y(l) and all points x(l)  >  2y(l) as $T<{{T}_{\text{c}}}$ since the domain wall fugacity flows to the y  =  0 line which corresponds to a completely ordered system with no spin flips. The high temperature phase is identified with the region of the x, y plane which flows to the region where $y(l)=\mathcal{O}(1)$ . This is interpreted as a system with a high concentration of domain walls corresponding to $T>{{T}_{\text{c}}}$ with spin disorder. Thus, one can identify ${{x}_{0}}-2{{y}_{0}}=-t$ so that

Equation (11)
Figure 3.

Figure 3. Flows from the linearized RG equations of (10).

Standard image High-resolution image

The unstable separatrix $C=0={{x}_{0}}+2{{y}_{0}}$ corresponding to t  =  4y0 or $T\gg {{T}_{\text{c}}}$ is just part of the disordered phase, see figure 3.

The solutions to the flow equations (10) fall into three distinct regions of the x(l), y(l) plane: region I, $C(t)\geqslant 0,\,\,\,\,x(l)\geqslant 2y(l)$ , region II with $C(t)<0,\,\,\,\,2y(l)>|x(l)|$ and region III where $C>0\,\,\,\,x(l)<0$ with $|x(l)|\geqslant 2y(l)$ . As will become apparent, region I corresponds to $T\leqslant {{T}_{\text{c}}}(\,y)$ , regions II and III correspond to $T>{{T}_{\text{c}}}(\,y)$ . In region I when $C(t)\geqslant 0$ and ${{x}_{0}}>2{{y}_{0}}$ , the solutions to (10) are

Equation (12)

From this, $y\left(l=\infty \right)=0$ so that the scaled system at $l=\infty $ has no defects which implies that, when $T\leqslant {{T}_{\text{c}}}$ , one can define a correlation length ${{\xi}_{-}}(T)={{\text{e}}^{{{l}^{\ast}}}}$ where ${{l}^{\ast}}=1/\left(2\sqrt{C}\right)$ , so that

Equation (13)

Thus, the true critical region is $t<{{t}^{\ast}}=4{{y}_{0}}$ where one may expect to see ${{\xi}_{-}}(T)$ diverging as ${{\text{e}}^{1/\left(4\sqrt{{{y}_{0}}|t|}\right)}}$ but this region is extremely small since the crossover temperature is proportional to the defect fugacity ${{y}_{0}}\sim {{\text{e}}^{-\mu /\left({{k}_{\text{B}}}T\right)}}$ which is exponentially small. Above this temperature but still in the region $|t|\ll 1$ , a gradual crossover will occur to the ${{\text{e}}^{1/\left(2|t|\right)}}$ behavior. This same phenomenon will be found in the 2D planar rotor model and related systems. The correlation length ${{\xi}_{-}}(T)$ can be interpreted as the mean separation between spin flips of opposite sign, or as the size of domains of reversed spins which diverges as $T\to T_{\text{c}}^{-}$ .

In this Ising ferromagnet with 1/r2 interactions, one can use the RG to show that the magnetization m(T) is discontinuous at ${{T}_{\text{c}}}$ [5]. From equation (7),

Equation (14)

This shows that K(l)m2(l) is a renormalization group invariant, at least to $\mathcal{O}\left(\,{{y}^{2}}\right)$ , so that one can obtain the value of the physical quantity K(0)m2(0) by a judicious choice of the length scale l. When $T\leqslant {{T}_{\text{c}}}$ , $y(\infty )=0$ so that the effective system at this scale contains no defects which implies that $m(\infty )={{m}_{\text{sat}}}=1$ and $K(\infty )=1+\sqrt{C}=1+\sqrt{|t|\left(4{{y}_{0}}\,+\,|t|\right)}$ so that

Equation (15)

Thus, the zero field magnetization m(T) has a discontinuous jump to $m\left(T>{{T}_{\text{c}}}\right)=0$ of size $K_{0}^{-1/2}\left({{T}_{\text{c}}}\right)=\sqrt{\frac{{{k}_{\text{B}}}{{T}_{\text{c}}}}{J}}$ , as first argued by Thouless [1], derived by the RG method [6, 8] and proved rigorously by Aizenman et al [17]. As mentioned above, the combination ${{K}_{0}}{{m}^{2}}$ plays the role played by the renormalized stiffness constant in the planar rotor model or in 4He in 2D as in figure 4.

Figure 4.

Figure 4. A plot of m2 as a function of T for various values of the fugacity y0. The straight dotted line is $m_{\text{c}}^{2}={{m}^{2}}\left(T_{\text{c}}^{-}\right)$ from (15).

Standard image High-resolution image

To this point, the discussion has been limited to region I where $y(l)\to 0$ as $l\to \infty $ so that $T\leqslant {{T}_{\text{c}}}$ . When $T>{{T}_{\text{c}}}$ but $t\leqslant 4{{y}_{0}}$ , the system is in region II. As the length scale l increases, the fugacity decreases but, at larger l, y(l) begins to increase and continues to increase as l does. This is readily seen from the linearized flow equations of (10) when C  <  0 and $4{{y}^{2}}-{{x}^{2}}=\,|C|>0$ , so that

Equation (16)

This is valid for any l, provided only that the RG equations (10) remain valid which requires that the renormalized fugacity y(l)  <  1. Thus, a suitable choice of l is l* where $x\left({{l}^{\ast}}\right)=-{{x}_{0}}$ , $y\left({{l}^{\ast}}\right)={{y}_{0}}$ with ${{x}_{0}}\gg \sqrt{|C|}$ and ${{y}_{0}}\gg \sqrt{|C|}$ , so that

Equation (17)

When t  >  4y0, the RG equations (10) must be used with $C\left(t,{{y}_{0}}\right)=t\left(t-4{{y}_{0}}\right)>0$ in region III. Most of this region correponds to systems with $T\gg {{T}_{\text{c}}}$ which is expected to display no critical behavior.

These results for the correlation lengths ${{\xi}_{\pm }}(t)$ have a physical interpretation for this 1D Ising model with 1/r2 interactions. In the low temperature region I, ${{\xi}_{-}}(T)$ may be interpreted as the mean separation of a pair of opposite domain walls or the size of a domain of reversed spins. In region II, the fugacity y(l) decreases from its initial value y0 to some minimum but increases for larger l to some value of $\mathcal{O}(1)$ which is outside the domain of validity of the renormalization group equations (7) and (10). Thus, choose l* in (17) so that $x\left({{l}^{\ast}}\right)=-{{x}_{0}}$ and $y\left({{l}^{\ast}}\right)={{y}_{0}}={{x}_{0}}+\mathcal{O}(C)$ so that $\text{ta}{{\text{n}}^{-1}}\left({{x}_{0}}/\sqrt{|C|}\right)=-\text{ta}{{\text{n}}^{-1}}\left(x\left({{l}^{\ast}}\right)/\sqrt{|C|}\right)=\pi /2$ and ${{\xi}_{+}}(T)$ is interpreted as the typical size of an ordered domain of spins when $T>{{T}_{\text{c}}}$ .

The singular part of the free energy fs is obtained from the flow equations

Equation (18)

The trick is to choose a value of l* such that the integral and f(l*) are easily obtained and a suitable choice is ${{l}^{\ast}}=\text{ln}{{\xi}_{\pm }}(T)$ . Thus the the singular part of the physical free energy fs  =  f(l  =  0) is given by

Equation (19)

This result was first obtained by an RG procedure [8]. This means that the specific heat ch(t) has the same essential singularity at ${{T}_{\text{c}}}$ as ${{\xi}_{\pm }}(T)$ which is undetectable since all derivatives of the specific heat are zero at ${{T}_{\text{c}}}$ [8].

2. Two dimensional superfluids and crystals

While I was still trying to understand Anderson's scaling papers, David Thouless drew my attention to some other problems which seemed to be closely related to the 1/r2 ferromagnetic Ising model, namely, the planar rotor model in 2D, films of 4Helium and melting of 2D crystals. These systems provided a major puzzle because of the argument of Peierls [12, 13] which excludes long range translational order in 2D. The mean square deviation $\langle {{\mathbf{u}}^{2}}\left(\mathbf{R}\right)\rangle $ of particles from their equilibrium positions $\mathbf{R}$ diverges with the system size L, $\langle {{\mathbf{u}}^{2}}\left(\mathbf{R}\right)\rangle \sim \text{ln}L$ . This means that there is no translational order in 2D. Using similar arguments, Mermin, Wagner and Wegner [2224] proved rigorously that there is no long range order in the 2D magnets with a continuous symmetry and short range interactions and there is no long range translational order in a 2D crystal. Hohenberg [25] proved that there is no Bose–Einstein condensation in 2D which apparently excludes superfluidity in 2D. At the time, the generally accepted conclusion was that there is no transition to a more ordered state at any finite temperature in a 2D system with a continuous symmetry and short range interactions. This excludes a low temperature ordered state in the 2D Heisenberg and XY magnets, in a thin film of 4He or nematic liquid crystal and in a 2D crystal, despite it being carefully pointed out that the theorem only excludes long range order in 2D but not a transition to a distinct low temperature phase without true long range order [22].

In the early 1970's the situation regarding transitions in these 2D systems was somewhat confused from both theoretical and experimental points of view. Numerical work on a system of hard disks in two dimensions [2631] indicated a transition between solid and fluid states which seemed to contradict the rigorous result [24] which forbids true long range order in a 2D crystal. The natural conclusion from this is that there is no transition from a crystal to a fluid in 2D which conflicts with the numerical study. A similar conflict arose between the rigorous result that long range order is forbidden in a 2D magnetic system with spins $\mathbf{s}=\left({{s}_{1}},{{s}_{2}},\cdots,{{s}_{n}}\right)$ where $n\geqslant 2$ interacting with short range interactions and series expansion results [3234] which indicated that a phase transition does occur. The evidence for a transition was rather inconclusive but was stronger for the 2D n  =  2 planar rotor model than for the n  =  3 Heisenberg model [34].

Some experimental work on 4Helium films indicated that there is a transition to a superfluid at ${{T}_{\text{c}}}>0$ with a discontinuous jump to zero of the superfluid density [3541] which was interpreted as a first order transition [42], despite the lack of a specific heat discontinuity [43]. The most convincing pieces of experimental evidence for a transition to a superfluid were experiments by Kagiwada [35] where the velocity of third sound, ${{c}_{3}}\propto \sqrt{{{\rho}_{s}}}$ was measured as shown in figure 5 (right) and and by Chester [36, 37] where the change in the resonant frequency of a quartz crystal due to the superfluid fraction of an adsorbed layer of 4He was studied as shown in figure 6. When we began to think about the transition in 2D superfluid films, the experimental situation could be summarized by figures 5 and 6. From these, we concluded that the areal superfluid density drops very rapidly and even might vanish discontinuously at ${{T}_{\text{c}}}$ in 2D.

Figure 5.

Figure 5. Left: Temperature dependence of areal superfluid density for films of several coverages. The vertical axis is the angular momentum at the same angular velocity of the film. Reprinted from [41] with permission. Copyright 1972 American Physical Society. Right: Third sound velocity ${{c}_{3}}\propto \sqrt{{{\sigma}_{s}}}$ against P0  −  P, which is a measure of the chemical potential μ for several fixed temperatures. The third sound signal vanishes at the hatched vertical lines indicating that ${{\sigma}_{s}}\left(T_{\text{c}}^{-}\right)>0$ . Reprinted from [35] with permission. Copyright 1973 American Physical Society.

Standard image High-resolution image
Figure 6.

Figure 6. Left: The horizontal axis measures the area mass density of the 4He film and the vertical axis is $-\Delta f$ , the reduction of the resonant frequency from the same film in the normal state. The deviation from the upper straight line is a measure of the superfluid mass density which decouples from the oscillating substrate. Reprinted from [36] with permission. Copyright 1969 American Physical Society. Right: The horizontal axis measures the total mass density σ adsorbed and the vertical axis the superfluid mass density ${{\sigma}_{s}}$ in units of 10−8 gm cm−2. Reprinted from [37] with permission. Copyright 1974 American Chemical Society.

Standard image High-resolution image

Thouless and I were faced with the task of constructing a plausible theory of superfluidity which would explain why the superfluid density was finite at $T_{\text{c}}^{-}$ and vanished discontinuously at $T_{\text{c}}^{+}$ . This theory would also have to accommodate that the specific heat peak near the transition is finite and smooth at ${{T}_{\text{c}}}$ [43] and the the low T susceptibility of the 2D planar rotor model is infinite [44] with zero magnetization. Even before we began to think about the problem, it was clear that the conventional wisdom about critical phenomena was not very useful and some out of the box thought was needed. This was exactly the style of problem which suited us: Thouless liked strange and unusual challenges and Kosterlitz was so ignorant of statistical mechanics and phase transitions that this problem was no different from any other so ideas of topological defects seemed difficult but normal. This was one of the rare situations when ignorance of conventional wisdom was a strength rather than a weakness.

2.1. Topological order

For both an equilibrium film of 4He and a 2D planar rotor magnet, the Hamiltonian is

Equation (20)

where a0 is the lattice spacing and

Here, J is the exchange interaction between nearest neighbor unit length spins, $\mathbf{s}\left(\mathbf{r}\right)=\left(\text{cos}\theta \left(\mathbf{r}\right),\text{sin}\theta \left(\mathbf{r}\right)\right)$ , $H\left[\left\{\mathbf{s}\right\}\right]=$ $\frac{J}{2}\sum_{<\mathbf{r},{{\mathbf{r}}^{\prime}}>}{{\left[\mathbf{s}\left(\mathbf{r}\right)-\mathbf{s}\left({{\mathbf{r}}^{\prime}}\right)\right]}^{2}}=J\sum_{<\mathbf{r},{{\mathbf{r}}^{\prime}}>}\left[1-\text{cos}\left(\theta \left(\mathbf{r}\right)-\theta \left({{\mathbf{r}}^{\prime}}\right)\right)\right]$ . At low T, adjacent spins will be almost parallel so that $\text{cos}(\delta \theta )\approx $ $1-(\delta \theta )/2$ giving (20). For a 4He film,

Equation (21)

where ${{\mathbf{v}}_{s}}\left(\mathbf{r}\right)$ is the local superfluid velocity relative to the substrate. Since ${{\mathbf{v}}_{s}}=\frac{\hslash}{m}\nabla \theta $ where θ is the phase of the condensate wave function $\psi \left(\mathbf{r}\right)=|\psi \left(\mathbf{r}\right)|{{\text{e}}^{\text{i}\theta \left(\mathbf{r}\right)}}$ . Thus, one argues that (20) is a good description of a planar rotor ferromagnet and of a thin 2D 4He film and this form is extensively used.

However, it is important to remember that this Hamiltonian is periodic under $\theta \left(\mathbf{r}\right)\to \theta \left(\mathbf{r}\right)+2\pi $ , which allows vortex excitations. Although these have large energy and are therefore improbable, vortices are the only excitations capable of disordering the system or, equivalently, of destroying a uniform superfluid flow. The drift of vortices across the flow to the edges of the system unwinds the phase difference of the superfluid order parameter $\psi ={{\text{e}}^{\text{i}\theta}}$ between the ends of the system. This simple observation implies that these topological excitations must be taken seriously and cannot be ignored because the are very improbable. Smooth fluctuations in the phase are the most probable excitations since these have arbitrarily small energy and are third sound modes in 4He films and spin waves in magnets but have no effect on the rigidity of the 2D system. These cannot have anything to do with the phase transition. Very similar but technically more complex arguments hold for a 2D crystal in which phonons are the low energy excitations which do not affect the transition to a fluid while the high energy topological excitations, dislocations, are responsible for the melting transition in 2D.

Note that the Hamiltonian of (20) can be obtained from a Ginsburg-Landau free energy density functional $\mathcal{F}\,[\psi ]$ [45]

Equation (22)

Since the system is assumed to be well below the mean field transition temperature, one take $r(T)\ll 0$ and $u\gg 0$ . The fluctuations in the amplitude $|\psi \left(\mathbf{r}\right)|$ are negligible and can be ignored, so that $|\psi {{|}^{2}}\,=\,|r(T)|/u$ which is a constant $\mathcal{O}(1)$ . Thus, in this phase only approximation, the Hamiltonian (20) results.

Ignoring effects of the $2\pi $ periodicity, at low T the obvious approximation is to regard the system as Gaussian with $-\infty <\theta \left(\mathbf{r}\right)<+\infty $ so that, with ${{s}_{x}}+\text{i}{{s}_{y}}={{\text{e}}^{\text{i}\theta}}=\psi $

Equation (23)

This result means that, as expected, there is no true long range order at any finite T  >  0 in the planar rotor magnet [23, 46, 47], no Bose Einstein condensation with true long range phase coherence in 2D 4He films [4850] and no long range order in 2D crystals [44, 51]. Also, the power law decay indicates that D  =  2 is the lower critical dimension.

The question of the lower critical dimension in models with a continuous symmetry with a Hamiltonian of the form

Equation (24)

was finally settled by Polyakov [52] by a renormalization group method. The RG recursion relation in dimension d for the temperature variable K−1 is

Equation (25)

This yields a critical fixed point at $1/{{K}^{\ast}}=2\pi (d-2)/(n-2)$ which vanishes as $d\to 2$ . In exactly d  =  2, the effective temperature K−1(l) increases with l implying that the system is in a disordered phase at any finite K−1  >  0. On the other hand, in exactly 2D, the n  =  2 model is marginal with $\text{d}{{K}^{-1}}/\text{d}l=0$ to all orders in K−1 [53]. The correlation function of (23) agrees with this result and is exact within the gaussian or spin wave approximation.

The same considerations apply to a 2D crystal whose ground state is generally a triangular lattice with sites denoted by $\mathbf{r}$ and deviations from these by $\mathbf{u}\left(\mathbf{r}\right)$ so that the elastic energy can be written in continuum notation as [54]

Equation (26)

Here, μ and λ are Lamé coefficients and the strain tensor ${{u}_{ij}}=\frac{1}{2}\left(\frac{\partial {{u}_{i}}}{\partial {{r}_{j}}}+\frac{\partial {{u}_{j}}}{\partial {{r}_{i}}}\right)$ . Ignoring the periodicity under $\mathbf{u}\to \mathbf{u}+\mathbf{a}$ , where $\mathbf{a}$ is a primitive lattice vector, one obtains the Debye Waller factor which is a measure of translational order. Defining the order parameter ${{\rho}_{\mathbf{G}}}\left(\mathbf{r}\right)={{\text{e}}^{\text{i}\mathbf{G}\centerdot \mathbf{r}}}={{\text{e}}^{\text{i}\mathbf{G}\centerdot \mathbf{u}\left(\mathbf{r}\right)}}$ with $\mathbf{G}$ a reciprocal lattice vector, one has

Equation (27)

This means that the long range translational order in the ground state is destroyed by the low energy phonon excitations at any T  >  0 in a 2D crystal, in complete analogy to spin waves in the 2D planar rotor ferromagnet.

The planar rotor and related models are special cases in which topological defects play a vital role [18, 19]. Minimizing H of (20), one obtains

Equation (28)

Because of the periodicity under $\theta \left(\mathbf{r}\right)\to \theta \left(\mathbf{r}\right)+2\pi n$ , the solution to (28) obeys

Equation (29)

where C encloses the vortex core of circulation n. It is easily seen that

Equation (30)

Here, $\Theta\left(\mathbf{r},\mathbf{R}\right)$ is the phase at $\mathbf{r}=(x,y)$ due to a unit strength vortex at $\mathbf{R}=(X,Y)$ on the dual lattice [19, 55, 56] corresponding to a local energy minimum, ni is the quantized circulation of the vortex at ${{\mathbf{R}}_{i}}$ and ϕ is the smooth deviation from the local minimum configuration, corresponding to spin waves in a magnet and to third sound in a 4He film. After some tedious but straghtforward algebra, one obtains an expression for the energy of a configuration of phase angles in the presence of an arbitrary set of vortices $n\left(\mathbf{R}\right)$ [19],

Equation (31)

Note that the spin wave and vortex contributions to H decouple because ${{\nabla}^{2}}\Theta\left(\mathbf{r}\right)=0$ .

This decoupling of the spin wave and vortex degrees of freedom is realized explicitly in the Villain model on a discrete lattice [55, 56] where the decoupling of the spin wave and the vortex parts is made explicit. The partition function is

Equation (32)

where the integer variable $m\left(\mathbf{r},{{\mathbf{r}}^{\prime}}\right)$ lies on the bond between nearest neighbor sites $\mathbf{r}$ and ${{\mathbf{r}}^{\prime}}$ . One can write $m\left(\mathbf{r},{{\mathbf{r}}^{\prime}}\right)=\tilde{m}\left(\mathbf{r},{{\mathbf{r}}^{\prime}}\right)+m\left(\mathbf{r}\right)-m\left({{\mathbf{r}}^{\prime}}\right)$ , by making a discrete gauge transformation, and then define the site variable as $\phi \left(\mathbf{r}\right)=\theta \left(\mathbf{r}\right)+2\pi m\left(\mathbf{r}\right)$ . This form of the interaction energy $V(\theta )$ is periodic under $\theta \left(\mathbf{r}\right)\to \theta \left(\mathbf{r}\right)+2\pi n\left(\mathbf{r}\right)$ and approximates ${{K}_{0}}\left(1-\text{cos}\theta \right)$ [55]. The difference between the Villain form and (31) is quantitative but not qualitative and, invoking the concept of universality [57], one expects all models with $2\pi $ periodicity and short range interactions to have identical behavior in the critical region.

One is now faced with computing the grand canonical partition function $Z\left({{K}_{0}},{{y}_{0}}\right)=\sum_{\left\{n\left(\mathbf{R}\right)\right\}}\text{exp}\left(-H/{{k}_{\text{B}}}T\right)$ . Here, L is the linear size of the system and a is a short distance cutoff which is the lattice spacing in a magnet and the vortex core size in a superfluid so that $L/a\gg 1$ . The last term in Hv of equation (31) requires that $\mathop{\int}^{}{{\text{d}}^{2}}\mathbf{R}n\left(\mathbf{R}\right)=0$ to ensure that the energy is finite in the thermodynamic limit $L\to \infty $ . The resulting expression is the energy of a neutral Coulomb plasma [1820] since the interaction between charged particles in 2D is ${{q}_{i}}{{q}_{j}}\text{ln}\left({{r}_{ij}}/a\right)$ and the problem is reduced to the statistical mechanics of the Coulomb plasma.

The parameter y0 in (31) is the fugacity of a vortex since the probability of a vortex of circulation $2\pi n$ is proportional to $y_{0}^{{{n}^{2}}}$ . To proceed, assume the vortex fugacity ${{y}_{0}}\ll 1$ so that the system has a very small concentration of vortices. This implies that one might try a perturbation expansion to $\mathcal{O}\left(\,y_{0}^{2}\right)$ where ${{y}_{0}}={{\text{e}}^{-{{E}_{\text{c}}}/{{k}_{\text{B}}}T}}$ . The vortex core energy ${{E}_{\text{c}}}$ is the interaction energy in the region near the vortex center where the quadratic approximation ${{(\nabla \theta )}^{2}}$ is inadequate because $\nabla \theta $ is large. This also defines the vortex core size a0 which is a few lattice spacings in a magnet and a few interparticle spacings in a superfluid. Near the center of a vortex, ${{v}_{s}}\sim 1/r$ which diverges as $r\to 0$ , which implies that the fluid is normal at the vortex core. Whatever the microscopic structure of a vortex core, there is a large energy density in the core region which is not described by ${{(\nabla \theta )}^{2}}$ and one assumes the contribution to the total energy from this region can be simply described by a core energy ${{E}_{\text{c}}}\gg {{k}_{\text{B}}}T$ . Furthermore, assume that the fugacity ${{y}_{0}}\ll 1$ and is a temperature independent constant. The description of the system by equation (31) is a very simplistic one containing only two sets of excitations which are the low energy spin waves and the disordering high energy vortices [1820]. All others are neglected and it is assumed that this description is adequate at low T and near ${{T}_{\text{c}}}$ since the transition is driven by the vortex excitations.

The first attempt [18] to solve the transition problem in the 2D planar rotor model was based on the free energy of a single isolated vortex of unit circulation. From (31), $\Delta E(L)/\left({{k}_{\text{B}}}T\right)=\pi {{K}_{0}}(T)\text{ln}\left(L/{{a}_{0}}\right)+\mathcal{O}(1)$ . The entropy is $\Delta S(L)/{{k}_{\text{B}}}=\text{ln}\left({{L}^{2}}/a_{0}^{2}\right)+\mathcal{O}(1)$ so that the free energy is $\Delta F(L)/$ $\left({{k}_{\text{B}}}T\right)=\left(\Delta E(L)-\Delta S(L)\right)/\left({{k}_{\text{B}}}T\right)=\left(\pi {{K}_{0}}(T)-2\right)\text{ln}\left(L/{{a}_{0}}\right)+\mathcal{O}(1)$ . Thus, in the thermodynamic limit $L/{{a}_{0}}\to \infty $ , the vortex free energy $\Delta F(L)\to -\infty $ when $\pi {{K}_{0}}(T)>2$ and $\Delta F(L)\to +\infty $ so that the probability of an isolated vortex being present in equilibrium is zero. On the other hand, when $\pi {{K}_{0}}(T)<2$ , $\Delta F(L)\to -\infty $ so that the probability of an isolated vortex in equilibrium is unity. Of course, there will be a non zero density of finite clusters of vortices of zero total vorticity at all temperatures since the free energy of a finite size zero vorticity cluster is finite. However, the transition temperature ${{T}_{\text{c}}}$ is determined by the renormalized stiffness ${{K}_{\text{R}}}(T)$ such that $\pi {{K}_{\text{R}}}\left({{T}_{\text{c}}}\right)=2$ , as discussed in the following section.

David and I congratulated ourselves on finding important new physics but our euphoria soon dissipated. We were informed that Berezinskii [47] had discussed the the vortex driven transition in a superfluid film a year earlier than our paper [18, 19]. Since neither of us knew any Russian we were blissfully unaware of this work while we were developing the basic physics of the vortex driven transition. For some unknown reason, our work seems to have had much greater impact than that of Berezinskii.

The careful reader will have realized that the discussion is rather sloppy and the relation between the integer field $n\left(\mathbf{R}\right)$ and the vorticity is not obvious. Also, the relation between the topological charges $q\left(\mathbf{R}\right)=1/(2\pi )\mathop{\oint}^{}\text{d}\theta =0,\pm 1$ and the effect of boundary conditions are unclear. These issues have been addressed in [58] where the relation between the topological charge $q\left(\mathbf{R}\right)$ and the integer variables $n\left(\mathbf{R}\right)$ is clarified. This issue was first raised by Savit [59]. The effects of boundary conditions in the Coulomb gas language is also discussed in [58] which is important for the Coulomb gas at very low temperature and for finding ground state configurations [6062].

2.2. Renormalization group for 2D Coulomb gas

In this section, I discuss the planar rotor magnet described by equation (31) at low T and fugacity ${{y}_{0}}\ll 1$ . The first attempt with David Thouless considered a neutral set of charges $n\left(\mathbf{R}\right)=\pm 1$ interacting by a 2D Coulomb potential (31). Since the fugacity ${{y}_{0}}\ll 1$ , charges with larger n are ignored because they are less probable. A neutral pair of unit charges separated by r interacting by a lnr Coulomb interaction will be screened by the polarization of smaller pairs separated by less than r which are screened by even smaller pairs. This leads to a scale dependent dielectric constant $\epsilon (r)={{K}_{0}}/K(r)$ where the force between the charges is $2\pi K(r)/r$ . The energy of this pair is

Equation (33)

We were unable to solve this integral equation so we made an unfortunate unnecessary approximation [19] by replacing $f\left({{r}^{\prime}}\right)$ by f(r) in (33). This was justified by $f\left({{r}^{\prime}}\right)-f(r)\ll 1$ , but led to incorrect results.

In an important paper [63], Young demonstrated that our incorrect approximation was unnecessary and presented the correct procedure by defining a scale dependent fugacity

Equation (34)

Differentiating these with respect to $l=\text{ln}r$ yields

Equation (35)

These are the recursion relations derived earlier [20] by a renormalization group method. However, even if Thouless and I had not made our incorrect approximation in our 1973 paper, but obtained the correct results of (35), we would not have known what to do with them as we knew nothing about scale dependent coupling constants or renormalization groups at that time.

I was not happy with the self consistent results in [19], and realized that the scaling methods of Anderson and coworkers [6, 8] for the 1D Ising model might be applicable to the vortex problem of equation (31). The thinking was that in both the 1D Ising ferromagnet in the previous section and the 2D planar rotor model, the interactions between the important topological point defects are logarithmic and the transitions might be very similar. Of course, in the early 1970's, the importance of the spatial dimension and the range of the interactions [57] was not known to me so, motivated by Anderson's solution of the Kondo problem [6, 8] the renormalization group equations for the 2D Coulomb gas were derived [20] by working with the partition function $Z\left({{K}_{0}},{{y}_{0}}\right)$ . This decouples into a product of a gaussian spin wave and a Coulomb gas partition functions,

Equation (36)

${{Z}_{\text{sw}}}\left({{K}_{0}}\right)$ is a trivial Gaussian partition function. In the Coulomb gas partition function ${{Z}_{\text{c}}}\left({{K}_{0}},{{y}_{0}}\right)$ , there are n q  =  +1 charges and n q  =  −1 charges. These correspond to vortices of $\pm 1$ circulation which interact by a 2D Coulomb interaction. The integrations over the vortex positions ${{\mathbf{R}}_{i}}$ are restricted by the hard core repulsion $|{{\mathbf{R}}_{i}}-{{\mathbf{R}}_{j}}|>a$ where the short distance cut-off length scale a can be taken to be a lattice spacing or the diameter of a vortex core. It turns out that this is quite arbitrary, but it must be kept finite until the end. The infrared cut-off is the system size L must also be kept finite during the calculations to avoid unphysical divergences. It is of some interest to note that both cut-offs, L and a, have clear physical meanings so there is no temptation to take the limit $L\to \infty $ or $a\to 0$ unless it is obvious that these limits make physical sense. Actually, these appear in the combination L/a and the real physical limit is $L/a\to \infty $ .

The original derivation of the RG flow equations was done by an extremely involved method [20] based on the method pioneered by Anderson and coworkers [6]. This involves integrating out the short distance degress of freedom in the partition function ${{Z}_{\text{c}}}\left({{K}_{0}},{{y}_{0}}\right)$ of equation (36) and rescaling the cut-off $a\to a{{\text{e}}^{l}}$ resulting in the RG flow equations

Equation (37)

where the lattice spacing at length scale l is $a(l)=a{{\text{e}}^{l}}$ . The interaction is K(l) and the effective fugacity is y(l). The reduced free energy density $f\left({{K}_{0}},{{y}_{0}}\right)$ and specific heat $c\left({{K}_{0}},{{y}_{0}}\right)$ are

Equation (38)

They can be integrated to give the renormalization group invariant

Equation (39)

The RG flows are shown in figure 7 for various values of C. The critical line ${{K}_{\text{c}}}(\,y)$ is the separatrix between regions I and II in figure 7 which is given by equation (39) with C(t, y0)  =  0. Since $y\ll 1$ ,

Equation (40)

and ${{K}_{\text{c}}}(\,y)$ from equations (39) and (40) is shown in figure 8

Figure 7.

Figure 7. RG flows for 2D planar rotor model from equation (39). Arrows denote the direction of flow for increasing l.

Standard image High-resolution image
Figure 8.

Figure 8. ${{K}_{\text{c}}}$ from equation (39) (solid line) and (40) (dashed line) which are valid only to $\mathcal{O}\left(\,y_{0}^{2}\right)$ .

Standard image High-resolution image

These recursion relations were later derived by a technically much simpler but more sophisticated method [56, 64] which starts with the Fourier transformed superfluid momentum density ${{\mathbf{g}}^{s}}\left(\mathbf{q}\right)$ correlation function

Equation (41)

Consider the renormalized stiffness constant proportional to the renormalized or measured superfluid density $\rho _{s}^{\text{R}}(T)$ , which can be written as a correlation function [65]

Equation (42)

Now, the vorticity–vorticity correlation function in (42) can be calculated as a power series in the fugacity y0 to leading order to obtain

Equation (43)

Since the renormalized stiffness constant ${{K}_{\text{R}}}$ is independent of the cut-off a, one can redefine the cut-off $a\to a{{\text{e}}^{l}}$ to obtain the renormalization group flows of equations (37) which is the same as the self consistent equation (33) which is correct to $\mathcal{O}\left(\,y_{0}^{2}\right)$ only. In principle, equation (43) can be calculated to any desired order. Since ${{K}_{\text{R}}}$ is independent of the cut-off, by redefining $a\to a{{\text{e}}^{l}}$ we can call K(l) and y(l) the effective coupling constat and fugacity at length scale l. This can be expressed as renormalization group equations (37) by standard methods.

This derivation has several advantages over the self consistent one as it makes clear that the RG equations (35) and (37) are valid to $\mathcal{O}\left(\,{{y}^{2}}\right)$ only. Also, this derivation also has the advantage that it uses the perturbative expansion of ${{K}_{\text{R}}}(K,y)$ which yields the exact relation [64]

Equation (44)

which is the Josephson relation [66] which, in d dimensions, is

Equation (45)

The first major physical prediction can be made from (37) and (44) at $T_{\text{c}}^{-}$ . From figure 8, one sees that $K\left(T_{\text{c}}^{-}\right)$ is an irrelevant variable which flows to the stable fixed point ${{K}_{\text{c}}}\left(l=\infty \right)=2/\pi $ and $y(\infty )=0$ . Thus, using (44), we have [64]

Equation (46)

This prediction has been checked experimentally [67, 68] and the data from several different experiments [6973] is presented in figure 9. It is of interest to notice that the experimental data was obtained and plotted before the authors were aware of the theoretical prediction of [64]. When the theoretical line was inserted, it fitted the plotted data to about 10% accuracy. This is actually a remarkable result because this theoretical prediction is a smoking gun prediction which is an inescapable consequence of the theory. As an aside, note that the stiffness constant ${{K}_{\text{R}}}(T)$ of (44) is sometimes called the helicity modulus $\Upsilon (T)$ [74, 75] which is related to ${{K}_{\text{R}}}(T)$ by

Equation (47)
Figure 9.

Figure 9. Results of third sound and torsional oscillator experiments for superfluid density discontinuity ${{\rho}_{s}}\left(T_{\text{c}}^{-}\right)$ as a function of temperature. Solid line is from equation (46) for the static theory. Reprinted from [68] with permission. Copyright 1978 American Physical Society.

Standard image High-resolution image

The RG equations (37) have a fixed line y(l)  =  0 which is stable for $\pi K(l)\geqslant 2$ and unstable for $\pi K(l)<2$ as shown in figure 7. Thus, one interprets region I of figure 7 as the low temperature phase $T<{{T}_{\text{c}}}$ , regions II and III as the disordered $T>{{T}_{\text{c}}}$ phase and the separatrix between regions I and II as the critical line $T={{T}_{\text{c}}}\left(\,{{y}_{0}}\right)$ . All points in the ${{K}_{0}}(T),{{y}_{0}}$ plane on the separatrix and in region I flow to the stable portion of the fixed line y  =  0, while all points in regions II and III flow to a region where the effective vortex fugacity y(l) is $\mathcal{O}(1)$ implying the system is disordered.

To proceed further, it is convenient to concentrate on the region near $y=0,\pi K=2$ by defining $\pi K(l)=2+x(l)$ and writing the flow equations (37) to lowest non-trivial order as [20, 56, 64]

Equation (48)

Integrating these leads to the renormalization group invariant

Equation (49)

The flows are plotted in figure 10 which is nothing but a linearized version of figure 7. Note that the renormalization group equations and consequences are very similar to those for the 1D long range Ising magnet [6, 8]. The line y  =  0 is identified as the spin wave phase since there are no free vortices when the fugacity y  =  0. This phase has finite rigidity and can be identified as the low T quasi ordered phase. On the other hand, when y  >  0 there will be small density of free, mobile vortices so that the rigidity vanishes at long length scales implying this is a high T disordered phase.

Figure 10.

Figure 10. Linearized flows in x, y plane from equation (49). The arrows denote the flow direction as l increases.

Standard image High-resolution image

The (x(l), y(l)) plane shown in figure 10 divides naturally into three separate regions; (I) $C\left(t,{{y}_{0}}\right)\geqslant 0,\,x(l)\geqslant 4\pi y(l)\geqslant 0$ , (II) $C\left(t,{{y}_{0}}\right)<0,\,|x(l)|<4\pi y(l)$ , and (III) $C\left(t,{{y}_{0}}\right)>0,\,x(l)<$ $-4\pi y(l)$ , which need separate solutions of equation (48). In region (I) $x(l)\geqslant 4\pi y(l)>0$ and $C\left(t,{{y}_{0}}\right)\geqslant 0$ , so that

Equation (50)

Now examine the behavior of x(l) in various limits by the deviation from ${{T}_{\text{c}}}$ small so that $C\left(t,{{y}_{0}}\right)\ll 1$ and the initial value ${{x}_{0}}\gg \sqrt{C}$ . From (50),

Equation (51)

This allows one to define a correlation length for $T\leqslant {{T}_{\text{c}}}$ by ${{\xi}_{-}}(T)={{\text{e}}^{{{l}^{\ast}}}}$ where l* is defined by $2{{l}^{\ast}}\sqrt{C}=1$ , which gives

Equation (52)

The renormalized, or measured, stiffness constant for $T\leqslant {{T}_{\text{c}}}$ is

Equation (53)

Equation (53) seems to give an experimentally measurable prediction for the superfluid density $\rho _{s}^{\text{R}}(T)$ shown in figure 11

Equation (54)
Figure 11.

Figure 11. The superfluid density ${{\rho}_{s}}(T)$ for different coverages showing the universal jump of ${{\rho}_{s}}\left(T_{\text{c}}^{-}/{{T}_{\text{c}}}\right)$ . Reprinted from [68] with permission. Copyright 1978 American Physical Society.

Standard image High-resolution image

However, the superfluid density $\rho _{s}^{\text{R}}(T)$ is, unfortunately, not accessible experimentally as measurements on 4He films have to performed with either, or both, the frequency $\omega \ne 0$ or the superfluid velocity ${{\mathbf{v}}_{s}}\ne 0$ . Of course, because the system is of finite size or because third sound waves are present $\mathbf{q}\ne 0$ , but this effect is not as important. The theoretical prediction here is the $q=\omega ={{\mathbf{v}}_{s}}=0$ component of the dynamical response function $\rho _{s}^{\text{R}}\left(q,\omega,{{\mathbf{v}}_{s}}\right)$ which, unfortunately, is inaccessible to experiment. In principle, one can deduce the behavior of the stiffness constant ${{K}_{\text{R}}}(T)$ in the whole region $T\leqslant {{T}_{\text{c}}}$ . For $T\ll {{T}_{\text{c}}}$ , ${{K}_{\text{R}}}(T)$ should decrease linearly as T increases until about $10\%$ below ${{T}_{\text{c}}}$ but then vortex fluctuations will take over and in the final approach to ${{T}_{\text{c}}}$ , the stiffness constant drops to the universal value $2/\pi $ as $\sqrt{|t|}$ [64]. The theoretical form of ${{\rho}_{s}}(T)$ is shown in figure 10 and an experimental measurement in figure 11. Note that the experimental ${{\rho}_{s}}(T)$ does not drop discontinuously to zero at ${{T}_{\text{c}}}$ but does drop rapidly. This is because the measurements [68] are performed at finite frequency while the theory [64] is for $\omega =0$ . A dynamical extension of the static theory is needed [76] to make a more convincing interpretation of the experiments on thin Helium films since the measurements are, of necessity, performed with at least one of $\left(\omega,{{\mathbf{v}}_{s}}\right)$ non zero, while the static theory is for $\omega =0={{\mathbf{v}}_{s}}$ only. The theoretical behavior of ${{\rho}_{s}}(T)$ as a function of temperature T is shown in figure 11 for different coverages.

In region II, C(t, y0)  <  0 in the small temperature region $0<t<8\pi {{y}_{0}}$ and x(l) can have either sign, so that from (50) [76]

Equation (55)

Here, the upper integration limit is chosen as $x\left({{l}_{+}}\right)=-{{x}_{0}}<0$ and, since the deviation from ${{T}_{\text{c}}}$ is very small when ${{x}_{0}}\gg \sqrt{|C\left(t,{{y}_{0}}\right)|}$ , so that y(l+)  =  y(0) when ${{l}_{+}}=\pi /\sqrt{|C|}$ . This allows one to define another length scale when t  >  0

Equation (56)

For length scales larger than ${{\xi}_{+}}$ , the effective fugacity y(l  >  l+) increases with l to large values outside the validity of our RG equations and the system is interpreted as completely disordered due to the large concentration of free unbound vortices.

The very reasonable assumption that ${{\rho}_{s}}(t>0)=0$ leads to the universal jump in ${{K}_{\text{R}}}(T)$ at ${{T}_{\text{c}}}$ . One can show that $\rho _{s}^{\text{R}}=0$ for $T>{{T}_{\text{c}}}$ by defining the generalized stiffness constant ${{K}_{\text{R}}}\left(q,{{K}_{0}},{{y}_{0}}\right)$ at finite q by, following Nelson [77],

Equation (57)

Equation (44) gives

Equation (58)

and we can evaluate this approximately at the scale ${{\text{e}}^{{{l}^{\ast}}}}={{\xi}_{+}}$ by a Debye–Hückel [78, 79] approximation

Equation (59)

where the average $\langle n\left(q{{\xi}_{+}}\right)n\left(-q{{\xi}_{+}}\right)\rangle $ must be calculated in the Coulomb gas ensemble with Hamiltonian

Equation (60)

At the scale ${{\text{e}}^{{{l}^{\ast}}}}={{\xi}_{+}}$ the system has large density of vortices with all integer values of ${{n}_{\text{r}}}$ , when a reasonable approximation is to integrate instead of summing over the n(q) to obtain

Equation (61)

Here, $B\left({{l}^{\ast}}\right)\approx -\text{ln}\,y\left({{l}^{\ast}}\right)=\mathcal{O}(1)$ , but its precise value is unimportant. Thus, one obtains

Equation (62)

as $q{{\xi}_{+}}\to 0$ . This means that ${{K}_{\text{R}}}(T)={{K}_{\text{R}}}\left(q=0,T,{{y}_{0}}\right)=0$ when $T>{{T}_{\text{c}}}$ as expected, which completes the argument for the universal jump ${{\rho}_{s}}\left(T_{\text{c}}^{-}\right)/{{T}_{\text{c}}}=3.491\times {{10}^{-9}}~\text{g}~\text{c}{{\text{m}}^{-2}}~{{\text{K}}^{-1}}$ [64].

In the context of the renormalization group, the length scale ${{\xi}_{-}}(t)$ is a measure of when the RG trajectory, which initially flows parallel to the critical separatrix, deviates significantly from the straight line. A more physical interpretation is that ${{\xi}_{-}}(t)$ for $t\leqslant 0$ is the maximum size of neutral bound vortex pairs and there are no vortex pairs with separation $r>{{\xi}_{-}}$ . Above ${{T}_{\text{c}}}$ , t  >  0 and ${{\xi}_{+}}(t)$ is the largest separation of bound vortex pairs. All pairs with separation $r>{{\xi}_{+}}(t)$ must be regarded as unbound free vortices which can move freely to dissipate superfluid flow or to disorder the system. It is important to understand that ${{\xi}_{+}}(t)<\infty $ for t  >  0 and ${{\xi}_{+}}(t)=\infty $ for $t\leqslant 0$ . On the other hand, the shorter length scale ${{\xi}_{-}}(t)<\infty $ both above and below ${{T}_{\text{c}}}$ and diverges as ${{\text{e}}^{\left(b/\sqrt{|t|}\right)}}$ as $|t|\to 0$ . It is important to note that both ${{\xi}_{+}}(t)$ and ${{\xi}_{-}}\left(|t|\right)$ diverge exponentially for $8\pi {{y}_{0}}>|t|\geqslant 0$ as (56)

Equation (63)

This behavior is commonly taken to define the critical region of the 2D planar rotor model and of a superfluid film but this region is limited to $|t|\ll 8\pi {{y}_{0}}$ and the vortex fugacity ${{y}_{0}}={{\text{e}}^{-{{E}_{\text{c}}}/{{k}_{\text{B}}}T}}$ is extremely small since the core energy ${{E}_{\text{c}}}\sim J$ . The real critical region should be taken to be the range of t for which ${{\xi}_{\pm }}\gg 1$ which is the region $\sqrt{|C(t,{{y}_{0}})|}\ll 1 $ where $|C\left(t,{{y}_{0}}\right)|\,=\,|t\left(8\pi {{y}_{0}}-t\right)|\ll 1$ in (63). There is a discussion of this crossover behavior of ${{\xi}_{\pm }}$ and the behavior outside the critical region in [80].

One can calculate other thermodynamic quantities from renormalization group considerations, particularly for $T\leqslant {{T}_{\text{c}}}$ when the RG maps the physical system into a simple Gaussian model when the vortex fugacity $y(l)\to 0$ . The two point correlation function G(r, K, y) was calculated by Kosterlitz [20] who obtained

Equation (64)

when $T={{T}_{\text{c}}}$ . I sketch the derivation [81] below by first noting the scaling equation for d  =  2,

Equation (65)

To obtain the physical correlation function $G(0)=G\left(r,{{K}_{0}},{{y}_{0}}\right)$ in terms of the physical parameters K0 and y0, one needs G(l) which is to be computed from the scaled Hamiltonian. The RG flow equations (48) are valid provided the fugacity y(l) remains sufficiently small, so l in (65) is chosen so that $r{{\text{e}}^{-l}}=1$ and G(1, K(l), y(l)) is needed. This is the correlation function $\langle \text{exp}\left[\text{i}\left(\theta \left(\mathbf{r}+\mathbf{e}\right)-\theta \left(\mathbf{r}\right)\right)\right]\rangle \approx 1$ because, at low T, the correlation function of nearest neighbor spins of unit length is unity. Thus, the singular part of the correlation function is

Equation (66)

To evaluate this, use $\eta (l)=1/4-x(l)/8+\mathcal{O}\left({{x}^{2}}(l)\right)$ with x(l) from equation (50) to obtain

Equation (67)

Thus, the correlation function for $T\leqslant {{T}_{\text{c}}}$ is

Equation (68)

where the temperature dependent exponent $\eta (T)$ is

Equation (69)

A plot of $\eta (T)$ is shown in figure 12 where one can see the $\sqrt{|t|}$ approach, for $|t|\ll 8\pi {{y}_{0}}$ , to the asymptotic value $\eta \left({{T}_{\text{c}}}\right)=1/4$ . Note that, for very large $r\gg {{\xi}_{-}}(T)$ , the correlation function has a pure power law decay with a temperature dependent exponent, $\eta (T)$ , which comes from an effective Gaussian model with coupling constant ${{K}_{\text{R}}}(T)$ . For $r\ll {{\xi}_{-}}(T)$ , $G\left(\mathbf{r},T\right)$ has the same power law decay but modified by a universal logarithmic correction in (68).

Figure 12.

Figure 12. The exponent $\eta (T)$ as a function of temperature from equation (69).

Standard image High-resolution image

The situation for $G\left(\mathbf{r},T\right)$ when $T>{{T}_{\text{c}}}$ is not so clear. Use equation (65) to obtain

Equation (70)

When $r\gg {{\xi}_{+}}(T)$ , choose $l=\text{ln}{{\xi}_{+}}$ but one still needs $G\left(r/{{\xi}_{+}},K\left(\text{ln}{{\xi}_{+}}\right),y\left(\text{ln}{{\xi}_{+}}\right)\right)$ which is expected to be $\mathcal{O}\left(\text{exp}\left(-r/{{\xi}_{+}}\right)\right)$ since the correlation function should decay exponentially. Thus, when $r\gg {{\xi}_{+}}(T)$ ,

Equation (71)

which was obtained explicitly by using a duality transformation [56]. On the other hand, when ${{\text{e}}^{l}}\ll {{\xi}_{+}}$ , even when $T>{{T}_{\text{c}}}$ , the correlation function $G\left(r/{{\xi}_{+}}\leqslant 1,K(l),y(l)\right)=\mathcal{O}(1)$ , so that

Equation (72)

where $l=\text{ln}r$ . The integral can be carried out by using $\eta (l)=1/4-x(l)/8$ , as before, but with x(l) from (55).

Equation (73)

Thus, we obtain

Equation (74)

When ${{\xi}_{+}}>r>{{\xi}_{-}}$ , $\pi >\sqrt{|C|}l>1/2$ where the behavior of the correlation function G(r) of (74) is not so clear. The factor in the brackets does not vary much for $\sqrt{|C|}l\sim \mathcal{O}(1)$ and the cross over to the behavior for $r\gg {{\xi}_{+}}$ is unclear.

The lowest order flow equations (37) yield a remarkable amount of information and appear to yield exact results for $T\leqslant {{T}_{\text{c}}}$ because y(l) is an irrelevant variable so that the system becomes Gaussian as $l\to \infty $ . However, one sees that this marginally irrelevant variable is responsible for the ${{\left(\text{ln}r\right)}^{1/8}}$ factor in the two point correlation function G(r). Do higher order terms in the RG flow equations give a noticeable correction to the expressions for G(r)? This calculation has been performed [82, 83] who derived flow equations to next order in y(l) and found the correlation function at ${{T}_{\text{c}}}$ as

Equation (75)

Note that these results imply the finite size scaling form for the susceptibility at criticality for a $L\times L$ system

Equation (76)

This has been verified to great accuracy by numerical simulations on large systems up to L  =  216 [84, 85].

I now turn to thermal properties, in particular the specific heat of the planar rotor model in 2D and of a thin film of 4He. In general, the specific heat is the most obvious quantity to measure since it usually displays a singularity, albeit a relatively weak one, at a phase transition. For example, the 2D Ising ferromagnet has a specific heat singularity of the form $c(T)\sim \text{ln}|t|$ while the susceptibility $\chi (T)\sim |t{{|}^{-1.75}}$ . However, a measurement of the specific heat requires no special knowledge of the particular system such as the form of the order parameter or of the ordering field. For example, for an antiferromagnet the uniform susceptibility, or the response to a uniform magnetic field, has no dramatic divergence at ${{T}_{\text{c}}}$ but just a minor kink. The dramatic effect is in the staggered susceptibility which is difficult to measure as it is the response to a staggered magnetic field conjugate to the staggered magnetization. The specific heat of a ferromagnet and an antiferromagnet are the same and are measured by the same methods which require no knowledge of the symmetry of the system or the nature of the order parameter. In general, if a transition exists, a specific heat measurement will find it, except in cases such as the systems being considered here.

One needs the free energy $f\left({{K}_{0}},{{y}_{0}}\right)$ from which the specific heat is obtained by differentiation with respect to K0 as in (38). Integrating the flow equation (37) for f(l) yields [86, 87]

Equation (77)

When $T\leqslant {{T}_{\text{c}}}$ , one can take the upper limit $l\to \infty $ because the RG equations (37) are valid for all $0\leqslant l\leqslant \infty $ and ${{\text{e}}^{-2l}}f\left(K(l),y(l)\right)\to 0$ so that

Equation (78)

When $T>{{T}_{\text{c}}}$ , the fugacity y(l) is relevant and increases so that the RG equations become invalid for l  >  l*, where l* is chosen judiciously [86, 87] and

Equation (79)

The idea is to use the RG flow equations (37) for $l\leqslant {{l}^{\ast}}$ and estimate $f\left(K\left({{l}^{\ast}}\right),y\left({{l}^{\ast}}\right)\right)$ by an improved Debye–Hückel approximation. Detailed discussions of the matching point and calculations of $f\left(K\left({{l}^{\ast}}\right),y\left({{l}^{\ast}}\right)\right)$ are found in [86, 87]. The result is shown in figure 13. At the transition at $K_{\text{c}}^{-1}$ , the specific heat has an unobservable essential singularity of the form $c(T)\sim \xi _{+}^{-2}(T)$ where ${{\xi}_{+}}(T)\sim {{\text{e}}^{b/\sqrt{|t|}}}$ at the critical point $K_{\text{c}}^{-1}$ with a non universal peak at much higher temperature. Since the specific heat is mainly due to the vortex degrees of freedom, one can understand the peak as being due to the entropy released by the unbinding of vortex pairs of decreasing size as the temperature increases. Since a very small density of very large pairs unbind at ${{T}_{\text{c}}}$ , the specific heat is very small there. The density of vortex pairs increases as their size decreases and these unbind at higher T with a corresponding larger entropy release and larger specific heat. The maximum will be when the largest density of the smallest vortex pairs unbind, which is estimated to be $40\%$ above ${{T}_{\text{c}}}$ as shown in figure 13 [87].

Figure 13.

Figure 13. Specific heat of 2D planar rotor model from theory. The full curve is from a renormalization group calculation and the dashed (dotted) lines are exact low (high) temperature results. The transition is at $K_{\text{c}}^{-1}\approx 1.33$ which is well below the rounded maximum at ${{K}^{-1}}\sim 1.7$ and the asymptotic behavior as ${{K}^{-1}}\to \infty $ is $c={{k}_{\text{B}}}/2$ . Reprinted from [87] with permission. Copyright 1981 American Physical Society.

Standard image High-resolution image

2.3. Roughening of crystal facets

Rather surprisingly, as pointed out by Chui and Weeks [88, 89], the 2D planar rotor model is intimately related to models for the roughening of facets of crystals in thermodynamic equilibrium as temperature is increased. We represent the particle positions of an ideal flat facet as a 2D lattice of sites labelled by $\mathbf{r}$ . Even if the equilibrium configuration of a particular crystal is a smooth planar facet which may appear in a crystal grown at a particular temperature, but it also may not because of competing out of equilibrium facets with faster growth rates. It is well known that a facet with screw dislocations terminating in it will grow very much faster than the ideal equilibrium facet. There is some hope that a quantum system such as a 4He crystal immersed in superfluid may equilibrate sufficiently fast to be seen.

The partition function of the 2D Coulomb gas representation of the planar rotor model of (31) is written as [56]

Equation (80)

with L the facet size. Consider two cases (i) when the vortex fugacity y0  =  1, use the identity [90]

to obtain the discrete Gaussian model [56, 88, 89, 93]

Equation (81)

This is interpreted as a model of the facet in which the height of the facet at $\mathbf{r}$ , relative to a reference plane, is $h\left(\mathbf{r}\right)$ and neighboring columns of atoms differ by an integer number of lattice spacings. The temperature of the roughening model is proportional to K0 while the temperature of the dual planar rotor is $K_{0}^{-1}$ . When K0 is small, it is clear that all columns tend to have the same height $h\left(\mathbf{r}\right)$ and the facet is smooth and flat. On the other hand, at higher temperature with K0 large, the column heights $h\left(\mathbf{r}\right)$ will vary with position leading to a rough facet with the correlation function $\langle {{\left(h\left(\mathbf{r}\right)-h(0)\right)}^{2}}\rangle \sim \text{ln}\,r$ .

(ii) ${{y}_{0}}\ll 1$ when the partition function becomes

Equation (82)

To convert this to the sine-Gordon roughening model [88, 89, 93], the height variable is defined by $h\left(\mathbf{r}\right)=b\phi \left(\mathbf{r}\right)$ where $h\left(\mathbf{r}\right)$ is the local height at $\mathbf{r}$ in steps of spacing b between lattice planes. The sine-Gordon representation of the roughening problem has the Hamiltonian

Equation (83)

where u  =  2y0 when ${{y}_{0}}\ll 1$ and a is the short distance cut off.

Flow equations for $\gamma (l)$ and u(l) [88, 93] are derived in a similar way to K(l) and y(l) in the vortex system by defining a renormalized stiffness ${{\gamma}_{\text{R}}}(T)$ as

Equation (84)

where $F\left(\mathbf{v}\right)$ is the free energy of a surface of area $\Omega $ and $\mathbf{v}$ is the average slope of the surface. Define $h\left(\mathbf{r}\right)=\mathbf{v}\centerdot \mathbf{r}+{{h}^{\prime}}\left(\mathbf{r}\right)$ with $\mathbf{v}=\left({{v}_{1}},{{v}_{2}}\right)$ so that the sine-Gordon Hamiltonian is

Equation (85)

Now expand $F\left(\mathbf{v}\right)=-{{k}_{\text{B}}}T\text{lnTr}\,\text{exp}\left[-H\left(\mathbf{v}\right)/{{k}_{\text{B}}}T\right]$ to $\mathcal{O}\left({{\mathbf{v}}^{2}}\right)$ to obtain

Equation (86)

where $K\equiv {{k}_{\text{B}}}T/\left(\gamma {{b}^{2}}\right)$ and we have used $\langle \text{cos}\left[(2\pi /b)\left({{h}^{\prime}}\left(\mathbf{r}\right)-\right. \right. $ $\left. \left. {{h}^{\prime}}(0)\right)\right]{{\rangle}_{0}}={{(r/a)}^{-2\pi K}}$ .

The last line of (86) yields renormalization group flow equations exactly as for the planar rotor model by rescaling the cut off $a\to a\,{{\text{e}}^{l}}$ ,

Equation (87)

This is of exactly the same form as the flow equations (37) except for the interpretation of the coupling constant K. For the planar rotor model, $K(T)={{(\hslash /m)}^{2}}\rho _{s}^{0}/\left({{k}_{\text{B}}}T\right)$ while, for the roughening model, $K(T)={{k}_{\text{B}}}T/\left(\gamma {{b}^{2}}\right)$ . The T dependence is inverted, as expected, because the models are duals of each other [56] and the low temperature phase of one is the high temperature phase of the other. Thus, in the high T phase of the roughening model, the coefficient u(l) of $\text{cos}\left(2\pi h\left(\mathbf{r}\right)\right)/b$ is irrelevant and the roughening Hamiltonian becomes a simple Gaussian model at long length scales with

Equation (88)

The integer spacing between the crystal planes becomes irrelevant and the interface becomes rough in the sense that the interface width $\langle {{h}^{2}}\left(\mathbf{r}\right)\rangle \sim \text{ln}L$ in a size L system.

One can immediately infer some exact results for the roughening model by exploiting the relationship with the planar rotor model. At the roughening transition, the universal relation of equation (46) becomes

Equation (89)

This follows directly from the planar rotor result and the duality relation and implies that facets with the largest lattice spacing b will have the highest roughening temperature ${{T}_{\text{R}}}$ , assuming the surface stiffness γ does not depend much on b. However, to convert these theoretical predictions into experimentally measurable quantities, one has to remember that the facets are the faces of a 3D real crystal, and the equilibrium theory of a constant volume crystal is required.

The basic idea is that the crystal shape is determined by minimizing the total surface free energy at fixed volume [91]. A flat facet with $-{{x}_{0}}<x<+{{x}_{0}}$ , in terms of reduced coordinates $\tilde{x}=x/L,\,\tilde{y}=y/L$ and $\tilde{h}=h/L$ , obeys

Equation (90)

Here, the reduced facet size ${{\tilde{\mathop{x}}\,}_{0}}\sim b/{{\xi}_{+}}(T)\sim \text{exp}\left(-B/\sqrt{\left(T-{{T}_{\text{R}}}\right)/{{T}_{\text{R}}}}\right)$ and ${{\tilde{h}}_{0}}={{f}_{0}}/\gamma $ , with f0 the surface free energy density.

For $T>{{T}_{\text{R}}}$ , the facet has vanished, ${{\tilde{x}}_{0}}=0$ , and the reduced height $\tilde{h}\left(\tilde{x},0\right)$ is

Equation (91)

The universal jump in the stiffness constant translates, for a crystal facet, to the reduced curvature of the rough surface ${{K}_{\text{R}}}(T)=(L/R){{k}_{\text{B}}}T/\left(\gamma {{b}^{2}}\right)=2/\pi $ at $T=T_{\text{R}}^{+}$ which jumps discontinuously to zero at $T_{\text{R}}^{-}$ . Here, L is the facet size and R is the radius of curvature [91, 93]. Measuring the predicted jump in the curvature of a disappearing facet is not easy and, to the best of my knowledge, not all the universal predictions have been verified experimentally, particularly that the facet size vanishes as $\xi _{+}^{-1}(T)$ [92, 94]. Attempts to measure the reduced curvature ${{K}_{\text{R}}}(T)$ of (89) have been made on hcp crystals in contact with superfluid [95], but with mixed success. Despite being the system most likely to be in equilibrium, the measurements are consistent with theory but suffer seriously from lack of equilibration.

2.4. Substrate effects

One motivation for developing a theory of transitions in 2D was the experimental results for very thin films of 4Helium adsorbed on a substrate [36, 37, 68]. This system is closely related to the 2D planar rotor model and the predictions from the theoretical model are directly applicable to the experiments provided one can argue that substrate effects are irrelevant in the renormalization group sense. Of course, the substrate is essential as a fluid film a few atomic lengths thick, a monolayer crystal or magnet cannot exist in the absence of a substrate. About the only systems which can exist without a substrate are suspended liquid crystal films [96103] and a monolayer of graphene [104]. Therefore, it is essential to assess the effects of all possible perturbations due to a substrate on the ideal two dimensional system of section 2.2.

Before discussing some of the substrate effects on 2D planar rotor systems, we can see why superfluidity in thin 4Helium films can be understood quantitatively by the idealized 2D theory which ignores the effects of the far from perfect Mylar substrate. This can be modeled by a random potential coupled to the local mass density $|\psi \left(\mathbf{r}\right){{|}^{2}}$ . The order parameter is the quantum mechanical condensate wave function $\psi \left(\mathbf{r}\right)=\,|\psi \left(\mathbf{r}\right)|{{\text{e}}^{\text{i}\theta \left(\mathbf{r}\right)}}$ and the random substrate potential cannot couple directly to the phase of the wave function but only to its amplitude and is irrelevant. This justifies using the ideal planar rotor model for a superfluid Helium film on a Mylar substrate.

Many other systems are influenced by the substrate particularly a planar rotor magnet in a symmetry breaking crystal field such as a planar ferromagnet on a periodic substrate described by a Hamiltonian

Equation (92)

where hp is the strength of the p-fold symmetry breaking field. When $|{{h}_{p}}|\gg 1$ , the free energy is minimized when $\text{cos}\theta =+1\,\,\Rightarrow \,\,\theta \left(\mathbf{r}\right)=2\pi n/p$ for hp  >  0, and, when hp  <  0, $\theta \left(\mathbf{r}\right)=(2n+1)\pi /p$ , with $n=0,1,\cdots,p-1$ . These form the p-state clock models which correspond to the Ising ferromagnet for p  =  2 and to the 3-state Potts model for p  =  3.

For $p\geqslant 5$ , there are many possibilities since higher harmonics of the site anisotropy are possible and also of the interaction terms. The most general Hamiltonian on a lattice with sites $\mathbf{r}$ with p-fold symmetry breaking is

Equation (93)

This Hamiltonian represents many statistical mechanical models in two dimensions such as Potts models [105], the Ashkin–Teller model [106], the 8-vertex model [107], etc. These can all be written as generalized Coulomb gases [108114]. For example, the 4-state Potts model is obtained when p  =  4, ${{h}_{4}}\to \infty $ , h8  =  0, so that $\theta \left(\mathbf{r}\right)=0,\pi /2,\pi,3\pi /2,$ and ${{K}_{1}}=2{{K}_{2}}$ . Then the Hamiltonian becomes $H/\left({{k}_{\text{B}}}T\right)=-{{K}_{2}}\left(2{{\delta}_{s{{s}^{\prime}}}}+1\right)$ with $s=0,1\cdots,3$ .

In the simplest case of n  =  1 only in (93), the symmetry breaking term in the partition function can be written as [55, 56]

Equation (94)

with ${{y}_{p}}\approx {{h}_{p}}/2$ . The Coulomb gas representation of the planar rotor in a p-fold symmetry breaking field becomes [56]

Equation (95)

The integer vortex charges $n\left(\mathbf{R}\right)$ lie on the sites $\mathbf{R}=(X,Y)$ of the dual lattice and the symmetry breaking charges $m\left(\mathbf{r}\right)$ lie on the original lattice with sites $\mathbf{r}=(x,y)$ . Neutrality conditions $\sum_{\mathbf{R}}n\left(\mathbf{R}\right)=0=\sum_{\mathbf{r}}m\left(\mathbf{r}\right)$ are satisfied by both types of charges. The two Coulomb gases are coupled by the angular term $ip\sum_{\mathbf{R},\mathbf{r}}n\left(\mathbf{R}\right)m\left(\mathbf{r}\right)\Theta\left(\mathbf{R},\mathbf{r}\right)$ . One can see immediately that equation (95) is self dual by relabeling $n\left(\mathbf{R}\right)\leftrightarrow m\left(\mathbf{r}\right)$ , ${{y}_{0}}\leftrightarrow {{y}_{p}}$ and ${{K}_{0}}\leftrightarrow {{p}^{2}}/\left(4{{\pi}^{2}}{{K}_{0}}\right)$ . The partition function is unchanged by these transformations [56], so that

Equation (96)

One can write down immediately the RG flow equations by rescaling the short distance cut off $a\to a{{\text{e}}^{l}}$ as expansions in the fugacities y0 and yp to lowest order [56]. These are consistent with the duality relation of equation (96)

Equation (97)

From these equations, when p  >  4, there is a region of temperature when both fugacities y0 and yp are irrelevant given by

Equation (98)

In this range of T, the system is in a massless Gaussian phase. At low temperature $K_{0}^{-1}<8\pi /{{p}^{2}}$ , yp(l) is relevant and the system is in the low T phase where there is long range order with the system in one of the p ordered states. The phase diagram is shown in figure 14 left for p  =  6 and for p  =  4 in figure 14 right. When $p\leqslant 3$ , there is no intermediate massless phase as one or both fugacities ${{y}_{0}},{{y}_{p}}$ are relevant for all K. This is interpreted as a direct transition from the large K ordered state to a low K disordered state. For p  =  2 and p  =  3, at ${{h}_{p}}=\infty $ it is easy to see that the Hamiltonian of (93) reduces to that of the Ising or the 3-state Potts model

Equation (99)
Figure 14.

Figure 14. Phase diagrams in the T, hp plane. Left: p  =  6 showing the intermediate phase. Right: p  =  4 with no intermediate phase but three critical lines with varying exponents. Asterisks denote lines and regions of continuously varying critical exponents. Reprinted from [56] with permission. Copyright 1977 American Physical Society.

Standard image High-resolution image

I have shown that the planar rotor model, Zp models and the discrete Gaussian model can be transformed into Coulomb gases which is useful for analyzing critical behavior. The transformation can be done exactly for Villain models [55, 108] because of the quadratic form of the Hamiltonian. However, many realistic models are not quadratic. For example, the solid on solid model in which the step energy is linear in $|h\left(\mathbf{r}\right)-h\left({{\mathbf{r}}^{\prime}}\right)|$ cannot be transformed simply into a Coulomb gas model. However, the Hamiltonian can be expanded in powers of $\nabla \theta $ , the leading term being quadratic. One can then perform a more standard renormalization group analysis by momentum shell integration [75] and all non gaussian terms with more than two gradients are irrelevant. These RG arguments imply that SOS etc models with other than Villain interactions differ from a Coulomb gas by irrelevant operators only. Thus, provided one assumes that these models and Coulomb gases can be connected by RG flows with no fixed points intervening, they are asymptotically equivalent. Given this qualitative assumption, one can make quantitatively exact conclusions [83]. For the special case, p  =  4, the multicritical point, $\pi K=2$ and ${{y}_{0}}={{y}_{4}}=0$ , is the meeting point of three critical lines of continuously varying exponents separating a high temperature disordered phase from two low T ordered phases with two different spin orientations with $\theta \left(\mathbf{r}\right)=n\pi /2$ or $\theta \left(\mathbf{r}\right)=(2n+1)\pi /4$ , where n  =  0, 1, 2, 3. This point is identical to the critical point of the F model [115], and the three critical lines meeting there are in the universality class of the Baxter 8-vertex model [116].

Other interesting models of experimental relevance are obtained by including both the first and second harmonics of the p-fold anisotropy. The planar rotor model with both four and eight fold anisotropy

Equation (100)

is an interesting theoretical model but can also describe the adsorption of hydrogen on the (1 0 0) surface of tungsten. When the hydrogen coverage is changed, h8 does not change but h4 can be driven through zero when a structural transition of the tungsten surface occurs [117, 118]. At low T, $\pi K\geqslant \pi /2$ , equation (100) has two ordered phases with finite magnetization in one of the four directions $\theta \left(\mathbf{r}\right)=n\pi /2$ with n  =  0, 1, 2, 3 when h4  >  0 and $\theta \left(\mathbf{r}\right)=(2n+1)\pi /4$ when h4  <  0. The other anisotropy term ${{h}_{8}}\text{cos}8\theta \left(\mathbf{r}\right)$ is irrelevant for ${{K}^{-1}}>\pi /8$ and relevant for ${{K}^{-1}}\leqslant \pi /8$ . In the temperature range $\pi /8\leqslant {{K}^{-1}}\leqslant \pi /2$ the ordered phases will be separated by a line of continuous transitions at h4  =  0 At lower temperatures, ${{K}^{-1}}<\pi /8$ , h8 becomes relevant which has important consequences. There are two possibilities depending on the sign of h8: (a) h8  >  0 when the orderings favored by h4 and by h8 are compatible, independent of the sign of h4. The transition between the two ordered phases at h4  =  0 will be discontinuous first order [117, 118]. In case (b) h8  <  0, the orderings favored by ${{h}_{4}},{{h}_{8}}$ are incompatible leading to an extra phase when $|{{h}_{4}}|$ is small with this phase separated from the other two ordered phases by continuous transitions in the 2D Ising universality class. Scenario (i) is compatible with experimental data [119, 120] which seems to favor a first order transition at low temperatures. Figure 15 illustrates schematic phase diagrams for h8  >  0 (left) and h8  <  0 (right).

Figure 15.

Figure 15. Left: Schematic phase diagram for the Hamiltonian of (100) in the ${{h}_{4}},{{K}^{-1}}$ plane with a fixed, compatible h8  >  0. For finite h4, the two transition lines for ${{K}^{-1}}>K_{\text{KT}}^{-1}$ are in the universality class of the planar rotor in a 4-fold symmetry breaking field. Between $K_{\text{KT}}^{-1}$ and the multi critical point $K_{m}^{-1}$ , there is a line of pure planar rotor transitions. At lower T, ${{K}^{-1}}<K_{m}^{-1}$ , the 8-fold symmetry breaking field h8 is relevant and there is a line of first order transitions indicated by the dashed line at h4  =  0. The lower bound is $K_{m}^{-1}\geqslant \pi /8$ from (97). Right: The phase diagram for a competing h8  <  0. This differs from (a) only for ${{K}^{-1}}<K_{m}^{-1}$ when h8 is relevant. The first order line opens up into two lines of Ising transitions, which terminate at $\pm 4|{{h}_{8}}|$ . Reprinted from [118] with permission. Copyright 1994 American Physical Society.

Standard image High-resolution image

Other systems where the first two harmonics have important effects are tilted hexatic phases of liquid crystals. Experiments on the transitions between different tilted hexatic phases in thermotropic liquid crystals have observed a weakly first order transition from a hexatic-I to a hexatic-F phase [121]. In an analogous layered lyotropic liquid crystal the ${{L}_{\beta I}}$ and ${{L}_{\beta F}}$ phases, which correspond to the hexatic-I and hexatic-F phases respectively, were seen to be separated by a third ${{L}_{\beta L}}$ phase with continuous Ising-like transitions between the phases as in figure 15(b) [122]. These observations can be explained by modeling the system in terms of coupled planar rotor models for the bond angle, $\theta \left(\mathbf{r}\right)$ , and the tilt azimuthal angle, $\phi \left(\mathbf{r}\right)$ , of the director with Hamiltonian [123125]

Equation (101)

which is very similar to (100). In the very low temperature regime of interest, but above the crystallization temperature ${{T}_{\text{m}}}$ , one can write (101) as

Equation (102)

A detailed analysis of this model system is carried out [125] with results agreeing with the experimental observations by adjusting the sign of h12 in (101). The phase diagram for this model is essentially identical to figure 15 with ${{h}_{4}}\to {{h}_{6}}$ and ${{h}_{8}}\to {{h}_{12}}$ [125].

Dierker, Pindak and Meyer found a remarkable five-armed star defect sketched in figure 16 in thin tilted hexatic liquid crystal films at low temperatures [126]. This can be described by the Hamiltonian of (102) with h12  >  0 so that there is a first order transition at h6  =  0 between a hexatic I phase for h6  >  0 and a hexatic F phase for h6  <  0 [125]. It is easy to see that the Hamiltonian is minimized by ${{\theta}_{-}}\left(\mathbf{r}\right)=2\pi n/6$ , $n=0,1,\cdots,5$ , so that ${{\theta}_{-}}\left(\mathbf{r}\right)$ can have a discontinuity of $2\pi /6$ . When a film is formed, it can have a vortex about which the tilt azimuthal angle $\phi \left(\mathbf{r}\right)$ goes through $2\pi $ . When the system is cooled into the hexatic I phase, the bond orientational order increases and $\theta \left(\mathbf{r}\right)$ becomes locked to $\phi \left(\mathbf{r}\right)$ by the relevant h6 term, so that the $2\pi $ vortex in $\phi \left(\mathbf{r}\right)$ becomes a $2\pi $ disclination in $\theta \left(\mathbf{r}\right)$ . To reduce its energy, this defect breaks up into an N armed star structure with energy

Equation (103)

where R is the length of an arm, ε is the energy per unit length in units of ${{k}_{\text{B}}}T$ . As shown in [125, 126], $\Delta E/\left({{k}_{\text{B}}}T\right)$ is obtained by solving Laplace's equation ${{\nabla}^{2}}{{\theta}_{+}}(r,\Theta)=0$ in the wedge $-\pi /N<\Theta<+\pi /N$ subject to the boundary conditions

Equation (104)

so that

Equation (105)
Figure 16.

Figure 16. The orientation of the director (arrows) and of the bond orientation (crosses) about the 5-armed defect at temperature (a) ${{T}_{\text{m}}}<T<{{T}_{I}}$ and (b) T  >  TI. Reprinted from [126] with permission. Copyright 1986 American Physical Society.

Standard image High-resolution image

For the equilibrium arm length, minimize this with respect to R so that [125]

Equation (106)

For reasonable physical values ${{K}_{6}}\gg {{K}_{1}},g$ , so that the parameters of (101) become $\alpha +\beta =1$ with $\alpha \gg \beta $ and $\gamma \sim {{10}^{3}}$ . For these parameters, the value N  =  5 minimizes the energy $\Delta E/{{k}_{\text{B}}}T$ , in agreement with observation [126]. N  =  6 would relieve completely the bond orientational strain between the arms but the cost of the extra arm is prohibitive [126].

The expression for the arm length R can be used to test experimentally one prediction of the KTHNY theory of 2D melting [125, 126]. The equilibrium arm length of the arms of the star defect is given by (106) and is proportional to the temperature dependent coupling constant ${{K}_{+}}(T)={{K}_{6}}+{{K}_{1}}+2g$ where K6 is the bond orientational stiffness called $K_{\text{A}}^{\text{R}}(T)$ in (153). In the liquid crystal system of [126], the coupling K+ (T) is dominated by the bond orientational stiffness ${{K}_{\text{A}}}(T)$ whose temperature dependence is sketched in figure 19. Although the experiment of [126] is probably not in the asymptotic regime, the data for the arm length R(T) close to the crystal-hexatic transition was fitted to the form [126]

Equation (107)

This yields $\tilde{\nu}=0.38\pm 0.15$ , which is consistent with the theoretical prediction $\tilde{\nu}=0.3696\cdots $ [127130]. This is one of the earliest experiments to test the KTHNY theory of melting and the observations are completely consistent with all the theoretical predictions.

There have been a number of investigations of freely suspended thin liquid crystal films [100, 131136] to look for the BKTHNY melting scenario and the elusive hexatic phase. This requires a material with the sequence of phases as T is lowered smectic-A—hexatic-B—crystal-B to avoid the extra complications due to tilted molecules. One of the aims was to investigate the predicted scaling relation between the harmonics of the bond orientational order parameter [137139]

Equation (108)

This has been verified in 3D [137] with $\lambda \approx 0.3$ and $\lambda =0.8$ for a thin smectic-C film [133, 134]. However, the theoretical value in 2D is $\lambda =1$ which follows from a low T Gaussian theory for the planar rotor model [138]. Finally, agreement was found between theory and experiment on a two layer film of smectic B 54COOBC which has bond but no herringbone order [139141]. Interestingly, early measurements claimed disagreement with KT theory because of a specific heat anomaly at the Sm-A—Hex-B transition but later work [141] found a new Sm-A' phase with a different density and that the specific heat anomaly is at the Sm-A—Sm-A' transition while the Sm-A'—Hex-B—Cry-B phase transitions are consistent with the disclination/dislocation scenario.

3. Melting of two dimensional crystals

In early papers [18, 19], Kosterlitz and Thouless suggested that a two dimensional crystal melts by the unbinding of bound pairs of dislocations. These are the appropriate topological defects which can destroy the translational order in a periodic crystal. The crystal was pictured as a harmonic elastic solid with a small density of dislocations imposed, analogous to the vortices imposed on the gaussian background in a 2D Helium film. It is now clear that these physical ideas are correct but incomplete [127129]. There are two types of order in a translationally periodic crystal: (i) translational order under $\mathbf{u}\left(\mathbf{r}\right)\to \mathbf{u}\left(\mathbf{r}\right)+\mathbf{a}$ where $\mathbf{u}\left(\mathbf{r}\right)$ is the displacement of the position of the particle whose equilibrium position is the lattice site $\mathbf{r}$ and $\mathbf{a}$ is a lattice vector, and (ii) orientational order of the crystal axes. An obvious but fundamental question is: what excitations are responsible for destroying these orders, in what order are they destroyed as T is raised and what is the nature of the transitions? The theory of melting in 2D was worked out in seminal papers in 1978–79 by Halperin, Nelson and Young (HNY) [128130] where they showed that, as T is raised, a 2D crystal melts to an anisotropic fluid at ${{T}_{\text{m}}}$ and, at higher $T={{T}_{I}}>{{T}_{\text{m}}}$ , there is a transition to the expected isotropic fluid phase with exponentially decaying translational and orientational order.

The breakthrough of HNY was to realize the importance of the angular terms in the dislocation-dislocation interaction. In the original work [18, 19], the angular terms were ignored on the grounds that they are subleading. By analyzing the system carefully, it was realized that the dislocation unbinding transition at ${{T}_{\text{m}}}$ did not yield an isotropic liquid, but an intermediate anisotropic fluid with exponentially decaying translational order and algebraically decaying orientational order, due to remnants of the long range orientational order of the low temperature crystal phase. At the higher temperature, TI, the orientational order is destroyed resulting in an isotropic fluid with exponentially decaying translational and orientational correlations. Some experimental and numerical investigations followed with little agreement with theory or with other numerical simulations [142]. Early experiments and simulations on 2D melting seemed to favor a discontinuous first order transition directly from a periodic solid to an isotropic fluid with no intermediate anisotropic fluid, in contradiction to theory [129]. Surprisingly, recent set of experiments on the melting of 2D colloidal crystals [143, 144] finally agree with theory but such experiments have been possible only over the last decade.

The notion of a periodic crystal in 2D was quite controversial in 1972 when Thouless and I began to think about the melting problem because of a rigorous theorem that there is no long range crystalline order in 2D [24]. This theorem, although correct, misled most physicists at that time because it proves exactly what it claims but does not exclude finite elastic moduli in 2D, as was understood by the author of the theorem. Thouless and I pictured a 2D crystal as a periodic array of points, representing the particles, on an elastic sheet. We asked what would happen to this array when the sheet is elastically deformed without tearing? It is obvious that the lattice does not change much locally, but the relative separation of a pair of initially distant points can undergo a very large change. The idea of a 2D crystal became much clearer with this pictorial representation. We realized that the essential defining characteristic of a 2D crystal is not long range translational order, which is destroyed by thermal fluctuations, but is the fact that the elastic moduli are finite.

This picture is written mathematically by writing the local number density $\rho \left(\mathbf{r}\right)$ of particles as

Equation (109)

where $\mathbf{r}$ is a site of the periodic reference lattice, $\mathbf{G}$ a reciprocal lattice vector such that $\mathbf{G}\centerdot \mathbf{r}=2\pi n$ and $\mathbf{u}\left(\mathbf{r}\right)$ is the displacement of the particle from its equilibrium position $\mathbf{r}$ . The most common crystal structure in 2D is the triangular lattice where each particle has six nearest neighbors. This structure also has a six-fold rotational symmetry as the crystal axes are oriented at angles $2\pi n/6$ relative to one another. The translational and orientational order parameters are [129]

Equation (110)

where $\theta \left(\mathbf{r}\right)$ is the angle the local crystal axis makes with some arbitrary reference direction. Note that ${{\rho}_{\mathbf{G}}}\left(\mathbf{r}\right)$ is invariant under $\mathbf{u}\left(\mathbf{r}\right)\to \mathbf{u}\left(\mathbf{r}\right)+\mathbf{a}$ with $\mathbf{a}$ a primitive lattice vector, and ${{\psi}_{6}}\left(\mathbf{r}\right)$ under $\theta \left(\mathbf{r}\right)\to \theta \left(\mathbf{r}\right)+2\pi /3$ .

A natural way to proceed is to write down a Ginsburg-Landau-Wilson free energy functional in terms of the order parameters ${{\rho}_{\mathbf{G}}}\left(\mathbf{r}\right)$ and ${{\psi}_{6}}\left(\mathbf{r}\right)$ which is invariant under the necessary symmetry transformations. This has been done by Halperin [145] following the treatment by de Gennes [146] with the free energy density [147]

Equation (111)

Here, consider only the six smallest reciprocal lattice vectors ${{\mathbf{G}}_{\alpha}}$ with $|{{\mathbf{G}}_{\alpha}}|\,={{G}_{0}},\,\,\alpha =1,2,\cdots,6,$ of the underlying triangular lattice and the corresponding Fourier components ${{\rho}_{\alpha}}={{\rho}_{{{\mathbf{G}}_{\alpha}}}}\left(\mathbf{r}\right)$ of the density. The unit vector $\mathbf{\hat{z}}$ is normal to the (x, y) plane of the crystal and $\theta \left(\mathbf{r}\right)$ is the phase of the bond-orientational order parameter

In a mean field approximation, the presence of the third order terms makes the transition first order [45]. ${{K}_{\text{A}}}$ is a bare rotational stiffness constant like that of the planar rotor model (20). The gradient terms are fixed by the transformation properties under rotations $\theta \left(\mathbf{r}\right)\to \theta \left(\mathbf{r}\right)+{{\theta}_{0}}$ which requires that

Equation (112)

so that the free energy (111) is invariant under rotations. Assume the transition occurs well below the mean field transition temperature so that $r(T)\ll 0$ and $|{{\rho}_{\alpha}}|={{\rho}_{0}}$ . In this phase only approximation [147],

Equation (113)

The free energy of (111) is minimized when

Equation (114)

and the elastic free energy of a 2D crystal becomes

Equation (115)

Here, the Lamé coefficients are given in terms of the constants A and B of (111) by [147]

Equation (116)

At this order, the elastic free energy is a quadratic functional of the linearized symmetric strain tensor

Equation (117)

In the same way as for the planar rotor model, the strain field can be decomposed as ${{u}_{ij}}\left(\mathbf{r}\right)={{\phi}_{ij}}\left(\mathbf{r}\right)+u_{ij}^{s}\left(\mathbf{r}\right)$ where ${{\phi}_{ij}}\left(\mathbf{r}\right)$ is the smooth part of the strain ${{u}_{ij}}\left(\mathbf{r}\right)$ and $u_{ij}^{s}\left(\mathbf{r}\right)$ is the singular part due to dislocations [148]. These are characterized by the value of the integral of the displacement $\mathbf{u}\left(\mathbf{r}\right)$ round a closed contour

Equation (118)

Here, $\mathbf{b}\left(\mathbf{r}\right)$ is the dimensionless Burger's vector, a0 is the crystal lattice spacing, ${{\mathbf{\hat{e}}}_{1,2}}$ are unit lattice vectors of the underlying triangular lattice and n, m are integers. Within continuum elasticity theory, one can show that [129]

Equation (119)

To solve the differential equation (119) for the Green's function ${{\tilde{G}}_{m}}\left(\mathbf{r},{{\mathbf{r}}^{\prime}}\right)$ , one considers a crystal with free boundary conditions [129] which needs von Neumann boundary conditions, $\nabla {{\tilde{G}}_{m}}\left(\mathbf{r},{{\mathbf{r}}^{\prime}}\right)=\text{constant}$ , when $\mathbf{r}$ is on the boundary, which gives

Equation (120)

The constant C  >  0 is a measure of the ratio of the dislocation core size to the short distance cut off a and can be absorbed into an effective core energy, as was done for a superfluid vortex. Physically, this dislocation core can be thought of as the region near the center of the dislocation where the displacement $\mathbf{u}\left(\mathbf{r}\right)$ is large so that linear elasticity breaks down. In exact analogy with the vortex fugacity y0, one absorbs this unknown scale dependent dislocation core energy into an analogous dislocation fugacity ${{y}_{0}}={{\text{e}}^{-{{E}_{\text{c}}}/{{k}_{\text{B}}}T}}$ . The end result is that the total elastic energy, ${{H}_{\text{E}}}$ , can be decomposed into a smooth part and a singular part due to the dislocations as [128, 129]

Equation (121)

where $\tilde{K}={{K}_{0}}a_{0}^{2}/\left({{k}_{\text{B}}}T\right)$ , $\tilde{\mu}=a_{0}^{2}\mu /\left({{k}_{\text{B}}}T\right)$ , and $\tilde{\lambda}=a_{0}^{2}\lambda /\left({{k}_{\text{B}}}T\right)$ . The dislocation density $\mathbf{b}\left(\mathbf{r}\right)$ obeys a neutrality condition $\sum_{\mathbf{r}}\mathbf{b}\left(\mathbf{r}\right)=0$ , because the energy of an isolated dislocation diverges as $\text{ln}\left(L/{{a}_{0}}\right)$ where L is the linear size of the system.

One can immediately conclude that a low temperature crystal phase will be one described by H0 of (121) with renormalized elastic constants ${{\tilde{\mu}}_{\text{R}}}(T)$ and ${{\tilde{\lambda}}_{\text{R}}}(T)$ which are renormalized downwards from their bare values by the presence of bound pairs and triplets of elementary dislocations [127130]. In this phase, there will be no free, unbound dislocations because the presence of mobile dislocations will cause the shear modulus $\tilde{\mu}(T)$ to vanish so that the system will have a fluid like response to a shear stress. This was worked out by these authors by a renormalization group treatment for melting of a 2D triangular crystal on a smooth substrate using equation (121) but treating the angular term carefully. At low T, the 2D crystal is described by the Gaussian Hamiltonian H0 of equation (121) which gives the Debye–Waller factor ${{C}_{\text{G}}}\left(\mathbf{r}\right)$ and the structure function $S\left(\mathbf{q}\right)$

Equation (122)

where ${{\langle \rangle}_{0}}$ means an average over H0. The algebraic decay of ${{C}_{\mathbf{G}}}\left(\mathbf{r}\right)$ means that there is no true long range translational order in 2D because the smooth phonon excitations are enough to make $\langle {{\rho}_{\text{G}}}\left(\mathbf{r}\right)\rangle =0$ , in agreement with the rigorous Mermin–Wagner theorem [22, 24], but the harmonic crystal has a finite measured shear modulus $\tilde{\mu}(T)>0$ .

Interestingly, the x-ray structure function S(q) diverges at small reciprocal lattice vectors $\mathbf{G}$ when ${{\eta}_{\text{G}}}(T)<2$ and has finite cusp singularities when ${{\eta}_{\text{G}}}(T)>2$ as sketched in figure 17 for $T\leqslant {{T}_{\text{m}}}$ . A more careful calculation of the structure function [149] gives the same power law decay but with some angular dependence

Equation (123)
Figure 17.

Figure 17. A schematic sketch of the structure function $S\left(\mathbf{q}\right)$ of a 2D crystal from equation (143). For $T\leqslant {{T}_{\text{m}}}$ , peaks for small $\mathbf{G}$ diverge as $|\mathbf{q}-\mathbf{G}{{|}^{-2+{{\eta}_{\text{G}}}}}$ but for larger $\mathbf{G}$ are finite cusps. For $T>{{T}_{\text{m}}}$ , all peaks are finite with a maximum height $\sim \xi _{+}^{2-{{\eta}_{\text{G}}}}$ . Reprinted from [238] with permission. Copyright 2002 Cambridge University Press.

Standard image High-resolution image

Unfortunately, this angle dependent behavior of the structure function is also beyond experimental capabilities so the universal predictions for the structure function are unlikely to be verified or falsified in the foreseeable future.

3.1. Renormalization group for melting

Dislocations [148] are assumed to be responsible for the melting of a 2D crystal [18, 19] and, in analogy with vortices in the 2D planar rotor model, we want a correlation function which gives the renormalized elastic constants. Nelson and Halperin [128, 129] argued that inverse of the tensor of renormalized elastic constants $C_{ijkl}^{\text{R}}$ is

Equation (124)

where P is the perimeter of the crystal and $\mathbf{\hat{n}}$ is a unit vector normal to the perimeter pointing outwards from the crystal. Note that this definition of the renormalized compliance tensor or inverse elastic tensor, $C_{\text{R};ijkl}^{-1}$ in terms of fluctuations at the perimeter is exactly how the compliances are measured. Applying Green's theorem to the expression for Uij in (124), one obtains [129]

Equation (125)

where $\mathbf{r}$ denote the sites of the triangular crystal lattice and $\mathbf{R}$ the sites of the hexagonal dual lattice on which live the dislocations. It is not hard to see that the dislocation contribution to $\mathop{\int}^{}{{\text{d}}^{2}}\mathbf{r}{{u}_{ij}}\left(\mathbf{r}\right)$ vanishes, so that the singular part of Uij is

Equation (126)

We have derived a suitable correlation function for the renormalized elastic constants, or response functions, from which we can deduce renormalization group flows, in exact analogy to the renormalized stiffness constant ${{K}_{\text{R}}}(T)={{\hslash}^{2}}\rho _{s}^{\text{R}}(T)/\left({{m}^{2}}{{k}_{\text{B}}}T\right)$ of equation (57). We have

Equation (127)

The correlation function $\langle U_{ij}^{s}U_{kl}^{s}\rangle $ is evaluated by a perturbation expansion in the dislocation fugacity $y={{\text{e}}^{-{{E}_{\text{c}}}/{{k}_{\text{B}}}T}}$ to obtain the RG flow equations [128130]

Equation (128)

Here, $\tilde{K}(l)$ is the coupling constant in the dislocation Hamiltonian ${{H}_{\text{D}}}/\left({{k}_{\text{B}}}T\right)$ of (121). The derivation of the flow equations (128) is identical in spirit to the vortex flows of (37) but more complex because there are three different elementary dislocations in a triangular lattice and only one elementary vortex in the planar rotor model. The origin of the term $\mathcal{O}\left(\,{{y}^{2}}\right)$ in the flow equation for the dislocation fugacity y(l) is that two different elementary dislocations can combine to form a third elementary dislocation [128130]. The RG flows for ${{\tilde{K}}^{-1}}(l)$ and y(l) are sketched in figure 18.

Figure 18.

Figure 18. RG flows in the x, y plane from equation (135) for b  =  0.4. The arrows denote the flow direction for increasing l.

Standard image High-resolution image

The recursion relations are very similar to those for the planar rotor model. Above a temperature ${{T}_{\text{m}}}$ , $\tilde{K}(l)<16\pi $ and the fugacity y(l) increases with l, implying that the dislocations unbind and become free to move under any small applied stress so that the elastic moduli vanish and the crystal has melted. The flow equations (128) up to a length scale ${{\xi}_{+}}(T)$ [129],

Equation (129)

with ${{b}^{\prime}}$ a non universal constant. The scale ${{\xi}_{+}}(T)$ can be interpreted as the scale below which dislocations are bound together in pairs or triplets of zero total Burger's vector. Equivalently, on length scales ${{\text{e}}^{{{l}^{\ast}}}}\leqslant {{\xi}_{+}}(T)$ , the system responds elastically, while on length scales ${{\text{e}}^{{{l}^{\ast}}}}>{{\xi}_{+}}(T)$ , the system responds like a fluid.

When $T\leqslant {{T}_{\text{m}}}$ , $\tilde{K}(l)\geqslant 16\pi $ and the fugacity $y(l)\to 0$ as $l\to \infty $ so that, at all length scales, the system responds elastically. The renormalization group transformation is derived from the perturbation expansion for the renormalized elastic constants ${{\tilde{\mu}}_{\text{R}}}(T)$ , ${{\tilde{\lambda}}_{\text{R}}}(T)$ and ${{\tilde{K}}_{\text{R}}}(T)$ so that [129]

Equation (130)

The Young's modulus, ${{\tilde{K}}_{\text{R}}}(T)$ is analogous to the renormalized stiffness ${{K}_{\text{R}}}(T,y)$ for a 4He film. At low T, corresponding to $\tilde{K}(l)\geqslant 16\pi $ , from (128), it is seen that the fugacity y(l) is irrelevant so that the system has no free dislocations at large length scales and the stiffness

Equation (131)

For $T\leqslant {{T}_{\text{m}}}$ , the values of ${{\tilde{\mu}}_{\text{R}}}(T)$ and ${{\tilde{\lambda}}_{\text{R}}}(T)$ can, in principle, be obtained by integrating (128) and they depend on the initial values ${{\mu}_{0}},{{\lambda}_{0}}$ and are non universal. However, the combination ${{\tilde{K}}_{\text{R}}}\left(T_{\text{m}}^{-}\right)=16\pi $ does have a universal value from (128), just like the universal stiffness of the 2D planar rotor model [129],

Equation (132)

To obtain more detailed behavior near the transition at $16\pi {{K}^{-1}}=1$ , assume the dislocation fugacity $y\ll 1$ , write $16\pi {{K}^{-1}}(l)=1+x(l)$ and expand the flow equations (128) to lowest order in x(l) to obtain [128130]

Equation (133)

The flows in the x, Y plane are obtained from

Equation (134)

which yields the renormalization group invariant

Equation (135)

Note that, when the parameter B  =  0, this reduces to the planar rotor problem since then (135) is just the square root of (49). The RG flows from (135) are shown in figure 18.

One can extract some more detailed behavior of the system in the vicinity of the melting transition at $T={{T}_{\text{m}}}$ by solving the recursion relations of equation (133) in detail [127, 129, 130]. To simplify the algebra, it is convenient to redefine the fugacity as $Y=\pi \sqrt{12A}y$ so that the recursion relations for x and Y become [130]

Equation (136)

On the critical line, Y(l)  =  −m0x(l) with ${{m}_{0}}=\sqrt{{{\alpha}^{2}}+2}-\alpha $ and

Equation (137)

For x0 slightly away from the critical separatrix, write [129, 130] Y(l)  =  −m0x(l)  +  D(l) where

Equation (138)

The initial value ${{D}_{0}}\propto t=\left(T-{{T}_{\text{m}}}\right)/{{T}_{\text{m}}}\ll 1$ . When t  <  0, the flow follows the critical separatrix for some distance and then Y(l) breaks away and plunges rapidly to for l  >  l*. One can calculate l* from the condition $D\left({{l}^{\ast}}\right)\approx Y\left({{l}^{\ast}}\right)$ which gives

Equation (139)

From this we can define the length scale ${{\xi}_{-}}(T)$ as the scale at which the deviation from the critical separatrix becomes significant [128130]

Equation (140)

This length scale exists both above and below ${{T}_{\text{m}}}$ and has the same interpretation as for the planar rotor model. When $T\leqslant {{T}_{\text{m}}}$ , ${{\xi}_{-}}(T)$ is the separation of the largest bound pair of primitive dislocations and there are no larger pairs. For $T>{{T}_{\text{m}}}$ , up to this scale, the dislocations can be regarded as bound in neutral pairs and triplets while at larger length scales the dislocations are unbound and freely mobile.

Since the dislocation fugacity y(l) scales to zero for $T\leqslant {{T}_{\text{m}}}$ , the system is elastic and the renormalized elastic constants ${{\tilde{\mu}}_{\text{R}}}(T)$ and ${{\tilde{\lambda}}_{\text{R}}}(T)$ are finite but non universal and are given by equation (131). The Young's modulus ${{\tilde{K}}_{\text{R}}}(T)$ has a universal value at $T_{m}^{-}$ of ${{\tilde{K}}_{\text{R}}}\left(T_{\text{m}}^{-}\right)=16\pi $ with [129]

Equation (141)

where c is a constant $\mathcal{O}(1)$ . This behavior has been verified in detail by experiment and simulation [144]. The x-ray structure function

Equation (142)

where the Debye Waller factor for $\mathbf{q}=\mathbf{G}$ is

Equation (143)

The structure function for $|\mathbf{q}-\mathbf{G}|\ll 1/a$ is the Fourier transform of the Debye–Waller factor ${{C}_{\mathbf{G}}}\left(\mathbf{r}\right)$ so that $S\left(\mathbf{q}\right)\sim |\mathbf{q}-\mathbf{G}{{|}^{{{\eta}_{\text{G}}}(T)-2}}$ . This diverges for the smaller values of $\mathbf{G}$ when ${{\eta}_{\text{G}}}<2$ as sketched in figure 17. Note that for a hexagonal lattice with lattice spacing a0, the smallest reciprocal lattice vector $|{{\mathbf{G}}_{0}}|\,=4\pi /\left({{a}_{0}}\sqrt{3}\right)$ so that

Equation (144)

For systems with long range repulsive interactions, such as electrons trapped above the surface of superfluid 4He [150] and colloidal particles at a water/air interface [144], the ratio ${{\tilde{\lambda}}_{\text{R}}}(T)/{{\tilde{\mu}}_{\text{R}}}(T)\gg 1$ so that

Equation (145)

This is not a very good estimate for a detailed comparison with theory because, the initial value of ${{\lambda}_{0}}=\infty $ is renormalized by dislocations to a finite value ${{\tilde{\lambda}}_{\text{R}}}\left(T_{\text{m}}^{-}\right)/{{\tilde{\mu}}_{\text{R}}}\left(T_{\text{m}}^{-}\right)\approx 10$ for colloidal particles interacting by a 1/r3 repulsive interaction [143] and is $\approx 23$ for electrons on 4He [151]. The quantity measured by experiment on the electron system is $\Gamma(T)=\sqrt{\pi {{n}_{s}}}{{e}^{2}}/{{k}_{\text{B}}}T=137\pm 15$ at $T={{T}_{\text{m}}}$ [150] and a remarkably good fit has been obtained by combining a molecular dynamics simulation with the renormalization group [151154].

Above ${{T}_{\text{m}}}$ , the dislocation fugacity is relevant and, just as in the planar rotor model, there is a correlation length ${{\xi}_{+}}(T)$ which is the length below which dislocations can be considered as bound and above which they are free [129, 130] in analogy with vortices in the planar rotor model [19, 20]. The positional correlation function will decay exponentially for $r\gg {{\xi}_{+}}$

Equation (146)

The scale dependent elastic constants ${{\tilde{\mu}}_{\text{R}}}(l),\,{{\tilde{\lambda}}_{\text{R}}}(l)=0$ for ${{\text{e}}^{l}}>{{\xi}_{+}}(T)$ but are finite for ${{\text{e}}^{l}}<{{\xi}_{+}}(T)$ . This is an example of a system which behaves as an elastic solid at short length scales ${{\text{e}}^{l}}<{{\xi}_{+}}$ and as a fluid for longer scales ${{\text{e}}^{l}}>{{\xi}_{+}}$ . The x-ray structure function

Equation (147)

as sketched in figure 17.

At first sight, this phase where dislocations are all unbound and free can be interpreted as a conventional isotropic fluid [18, 19, 127] as the translational order decays exponentially with distance. However, in their seminal work, Halperin and Nelson [128, 129] pointed out that this fluid phase is not the expected isotropic fluid but has remnants of the six-fold orientational order of the underlying hexagonal lattice of the low temperature solid phase. These possible orientational correlations may be described by a finite renormalized Frank constant ${{K}_{\text{A}}}(T)$ of the anisotropic fluid

Equation (148)

where θ is the angle of a bond between nearest neighbor particles relative to an arbitrary reference direction.

One expects that the distribution of bond angles $\theta \left(\mathbf{r}\right)$ are controlled by a free energy [128, 129],

Equation (149)

where ${{K}_{\text{A}}}(T)$ must be finite for the existence of orientational correlations. In the underlying solid phase, the bond angle ${{\theta}_{s}}\left(\mathbf{r}\right)$ due to a set of dislocations is

Equation (150)

from which one can write the stiffness constant as a correlation function,

Equation (151)

The dislocation correlation function is to be calculated from the dislocation free energy of equation (121) written in Fourier space as [128, 129]

Equation (152)

At this point, one sees why the sign of the angular term in the dislocation interaction in (121) is so important and why ignoring the apparently shorter ranged angular term leads to incorrect results [19, 127]. The existence of a finite Franck constant, ${{K}_{\text{A}}}(T)$ , depends on the transverse projection operator, ${{T}_{ij}}(q)={{\delta}_{ij}}-{{q}_{i}}{{q}_{j}}/{{q}^{2}}$ , in (152). One can understand the behavior of the orientational stiffness constant ${{K}_{\text{A}}}(T)$ by using the recursion relations (128) up to the limit of their validity, ${{l}^{\ast}}=\text{ln}{{\xi}_{+}}(T)$ , to obtain

Equation (153)

The scaling factor ${{\text{e}}^{2l}}$ arises from the two extra factors of q in the definition of ${{K}_{\text{A}}}$ in equation (151) compared to $\tilde{\mu}$ and $\tilde{\lambda}$ . At length scale ${{\text{e}}^{{{l}^{\ast}}}}={{\xi}_{+}}(T)$ , it is reasonable to assume that there is a large density of free dislocations in the system and a sensible approximation is to regard the dislocation density $\mathbf{b}\left(\mathbf{r}\right)$ as a continuous variable as all values of $\mathbf{b}\left(\mathbf{r}\right)$ are present at scale l*. This Debye–Hückel approximation allows the correlation function $\langle {{b}_{i}}\left(\mathbf{q}\right){{b}_{j}}\left(-\mathbf{q}\right)\rangle $ in (151) to be evaluated because the transverse part of the dislocation interaction in equation (152) does not contribute. One obtains

Equation (154)

which establishes that the fluid caused by free dislocations above ${{T}_{\text{m}}}$ does have orientational stiffness [128, 129] and is $not$ isotropic as assumed by Kosterlitz and Thouless [19].

Now one is left with a problem of a system with short range exponentially decaying translational order and algebraic orientational order described by phenomenological free energy of equation (149) with a finite bare orientational stiffness constant ${{K}_{\text{A}}}(T)$ of (154). This is the planar rotor problem discussed earlier in detail, the only difference that the topological defects are $\pi /3$ disclinations which interact logarithmically like vortices. The bond angle field $\theta \left(\mathbf{r}\right)$ is decomposed into a smooth part $\phi \left(\mathbf{r}\right)$ and a singular part ${{\theta}_{s}}\left(\mathbf{r}\right)$ which obeys

Equation (155)

where m is the total disinclination strength enclosed by the contour C. Thus one obtains the disclination free energy [128, 129] which is identical to the planar rotor problem of equation (31)

Equation (156)

Here $m\left(\mathbf{r}\right)=0,\pm 1,\cdots $ is a measure of the strength of the disinclination at $\mathbf{r}$ , a is the core size of a disinclination, ${{E}_{\text{c}}}$ is the disclination core energy and the disclinations are subject to the neutrality condition, $\sum_{\mathbf{r}}m\left(\mathbf{r}\right)=0$ .

Halperin and Nelson [128, 129] pointed out that one can immediately take over all the results for the superfluid film and showed there is a disinclination unbinding transition at T  =  TI to the high temperature isotropic fluid. The renormalized orientational stiffness constant just below TI is given by

Equation (157)

The renormalized Franck constant, $K_{\text{A}}^{\text{R}}(T)$ , behaves precisely like the renormalized stiffness constant of a 2D superfluid film except that, slightly above the melting temperature,

Equation (158)

When T  >  TI, the orientational order is short ranged,

Equation (159)

where the orientational correlation length ${{\xi}_{6}}(T)$ diverges exponentially as $T\to T_{I}^{+}$ ,

Equation (160)

Assuming that the initial value of the bare stiffness ${{K}_{0A}}\left({{l}^{\ast}}\right)$ is larger than the critical value $K_{0A}^{c}$ , the renormalized value of the orientational stiffness will flow to a value $K_{\text{A}}^{\text{R}}(T)\geqslant 72/\pi $ . For T slightly above the melting temperature ${{T}_{\text{m}}}$ , the theory predicts that $K_{\text{A}}^{\text{R}}(T)\sim {{\text{e}}^{2{{l}^{\ast}}}}=\xi _{+}^{2}(T)$ and, at $T_{m}^{-}$ , $K_{\text{A}}^{\text{R}}\left(T_{I}^{-}\right)=72/\pi $ [129]. Experimental measurements of $K_{\text{A}}^{\text{R}}(T)$ [158] agree well with these theoretical predictions. If, for some reason, the dislocation core energy ${{E}_{\text{c}}}$ of equation (154) is very small, it is possible that $K_{\text{A}}^{0}\left({{l}^{\ast}}\right)<{{K}_{0c}}$ and will flow immediately to zero. A hexatic phase will not exist, but this seems unlikely. It is also possible that the hexatic phase exists only over a very small range of ${{T}_{\text{m}}}<T\leqslant {{T}_{I}}$ . Below ${{T}_{\text{m}}}$ , the translational correlation length ${{\xi}_{+}}(T)=\infty $ , so that $K_{\text{A}}^{\text{R}}(T)=\infty $ , and there is true long range orientational order, as was first pointed out by Mermin [24]. The behavior of the renormalised orientational stiffness ${{K}_{\text{A}}}(T)$ is shown in figure 19.

Figure 19.

Figure 19. A schematic plot of the orientational stiffness constant $K_{\text{A}}^{\text{R}}(T)$ as a function of temperature. Reprinted from [238] with permission. Copyright 2002 Cambridge University Press.

Standard image High-resolution image

The most important, unexpected, and controversial prediction of the defect theory of melting is that, in 2D, the melting process is not a first order transition from a periodic solid to an isotropic fluid but proceeds by two successive continuous transitions with an orientationally ordered hexatic fluid between the low temperature periodic crystal and the high temperature isotropic fluid. In fact, this scenario has been verified quantitatively by careful experiments on a monolayer of paramagnetic colloidal particles trapped at a water/air interface [155158]. A recent very large scale simulation of hard disks [159161] verified the two step melting scenario with an intermediate hexatic phase with a continuous melting transition but a first order hexatic to isotropic fluid transition. However, in a very recent study [162] in the presence of pinning disorder, no signs of any first order character at either transition was observed. Since the original theoretical prediction of continuous two stage melting [129], there have been a number of experimental and numerical studies performed to test the heretical prediction of continuous melting by two continuous transitions with a hexatic fluid intervening between the solid and the isotropic fluid. At the time, conventional wisdom was that melting is a first order transition just as in 3D. The theoretical behavior of the orientational stiffness ${{K}_{\text{A}}}(T)$ is sketched in figure 19 which agrees fairly well with experiment.

3.2. Substrate effects on melting

Most experiments on 2D melting are performed on monolayer of adsorbed molecules on a supporting substrate such as graphite which has relatively large well oriented domains of binding sites arranged in a periodic array. One must therefore investigate the effect of such a periodic substrate on the possible states of the over layer and on the transitions from the most ordered to the most disordered states of the combined system. Halperin and Nelson investigated this in their seminal papers [128, 129], and the results are summarized here. Both the graphite substrate and the adsorbate overlayer have the same six-fold orientational symmetry but a generic substrate has a different periodicity to the natural periodicity of the adsorbed overlayer. This leads to the elastic free energy [128, 129]

Equation (161)

One has assumed that the substrate potential is weak and can be written as $V\left(\mathbf{r}\right)=\sum_{\mathbf{K}}{{h}_{\mathbf{K}}}{{\text{e}}^{\text{i}\mathbf{K}\centerdot \left(\mathbf{r}+\mathbf{u}\left(\mathbf{r}\right)\right)}}$ where $\left\{\mathbf{K}\right\}$ are the reciprocal lattice vectors of the periodic substrate potential. The $\left\{\mathbf{K}\right\}$ are assumed incommensurate with the reciprocal lattice vectors $\left\{\mathbf{G}\right\}$ of the adsorbate, assumed to be in a floating solid phase. This can be described by a harmonic free energy with renormalized elastic constants ${{\mu}_{\text{R}}}$ and ${{\lambda}_{\text{R}}}$ . Redefine the displacement field [129]

Equation (162)

where $\mathbf{G}$ is a reciprocal lattice vector of the adsorbed layer. The second term of equation (162) can be rewritten as

Equation (163)

where ${{\widehat{\boldsymbol{\epsilon }}}_{s}}\left(\mathbf{K}\right)$ is the ${{s}^{\text{th}}}$ polarization vector and ${{\omega}_{s}}\left(\mathbf{K}\right)$ is the eigenfrequency of the matrix ${{D}_{ij}}\left(\mathbf{K}\right)$ .

For a hexagonal over layer with lattice vectors $\left\{\mathbf{G}\right\}$ on a hexagonal substrate with incommensurate lattice vectors $\left\{\mathbf{K}\right\}$ , perfect alignment of the crystal axes with the substrate corresponds to $\theta =0$ , and one can write the Fourier expansion of $C(\theta )$ as

Equation (164)

and, for a hexagonal adsorbate on a square substrate

Equation (165)

Assuming that the relative orientations of the substrate and adsorbed layer vary slowly in space, θ is the bond angle $\theta \left(\mathbf{r}\right)=\frac{1}{2}\left(\partial u_{x}^{\prime}/\partial y-\partial u_{y}^{\prime}/\partial x\right)$ so that, for small deviations θ from perfect alignment, one obtains the effective elastic Hamiltonian [129, 130]

Equation (166)

To translate the coupling constants K1 and K2 into the language of [130] where the recursion relations were first worked out for this general dislocation Hamiltonian, one has

Equation (167)

From the recursion relations of [130], it follows that

Equation (168)

From this, it follows that dislocation unbinding melting takes place when the fugacity ${{y}_{0}}={{\text{e}}^{-{{E}_{\text{c}}}/{{k}_{\text{B}}}T}}$ becomes relevant, which is controlled by K1. The renormalized stiffness coefficient at $T_{\text{m}}^{-}$ is $K_{1}^{\text{R}}\left(T_{\text{m}}^{-}\right)=16\pi $ and the value of the exponent $\tilde{\nu}$ depends on $K_{2}^{\text{R}}\left(T_{M}^{-}\right)/K_{1}^{\text{R}}\left(T_{M}^{-}\right)$ . From (168) it is clear that K2(l) flows to some non-universal finite value at $l=\infty $ and, when K2  =  0, $\tilde{\nu}=2/5$ as found by Nelson [127] and melting on an incommensurate substrate is very similar to a smooth substrate. As pointed out by Nelson [77], the situation above ${{T}_{\text{m}}}$ is rather different. The orientational bias can be modeled by adding a term ${{h}_{p}}\text{cos}p\theta $ to the hexatic free energy

Equation (169)

Clearly, for a hexagonal overlayer with orientational order parameter ${{\psi}_{6}}={{\text{e}}^{6\text{i}\theta}}$ on a hexagonal substrate with p  =  6, h6 acts like a uniform magnetic field inducing long range orientational order for all values of $T>{{T}_{\text{m}}}$ , wiping out the hexatic to isotropic fluid transition on a smooth substrate at TI.

When the substrate and adsorbed layer reciprocal lattice vectors $\left\{\mathbf{G}\right\}$ and $\left\{\mathbf{K}\right\}$ have a set $\left\{\mathbf{M}\right\}$ in common, the over layer may become locked to the substrate as a commensurate phase. This situation can be studied by considering the smallest common reciprocal lattice vector of length ${{M}_{0}}=|{{\mathbf{M}}_{0}}|$ and writing the free energy as [129]

Equation (170)

where it is assumed that the effects of all incommensurate lattice vectors have been incorporated into the elastic part ${{H}_{\text{E}}}$ . This can be done because these and terms in the substrate potential

Equation (171)

with $|\mathbf{K}|>{{M}_{0}}$ are less relevant in the renormalization group sense than those with $|\mathbf{K}|={{M}_{0}}$ .

The relevance of the ${{h}_{\mathbf{M}}}$ is obtained from the correlation function computed from the renormalized elastic free energy ${{H}_{\text{E}}}/\left({{k}_{\text{B}}}T\right)$ with ${{h}_{\mathbf{M}}}=0$ ,

Equation (172)

From this, the renormalization group eigenvalue of ${{h}_{{{M}_{0}}}}(l)={{h}_{{{M}_{0}}}}(0){{\text{e}}^{{{\lambda}_{{{M}_{0}}}}l}}$ is

Equation (173)

From (173), at sufficiently low T, all the ${{h}_{\mathbf{M}}}$ are relevant so that the over layer is locked to the substrate, implying that a lattice gas description is more appropriate [163]. If the substrate is sufficiently coarse, ${{\lambda}_{{{M}_{0}}}}\left({{T}_{\text{m}}}\right)>0$ , so that the adsorbate is locked to the substrate up to the melting temperature ${{T}_{\text{m}}}$ , for ${{h}_{{{\mathbf{M}}_{0}}}}\to 0$ so that dislocation melting is inappropriate at any T. On the other hand, for a fine substrate mesh, M0 is large and ${{\lambda}_{{{M}_{0}}}}\left({{T}_{\text{m}}}\right)<0$ , so that the overlayer is unlocked from the substrate in a temperature range ${{T}_{l}}<T<{{T}_{\text{m}}}$ , leading to a floating incommensurate solid which may be accidentally commensurate with the substrate. Halperin and Nelson [129] have derived a criterion for the unlocking temperature ${{T}_{\text{L}}}<{{T}_{\text{m}}}$ . The consequence of this is that ideal 2D melting is still possible on a periodic crystal substrate provided the adsorbate is a floating incommensurate solid.

The most dramatic effect of the substrate potential is on the hexatic fluid phase $T>{{T}_{\text{m}}}$ . One performs an analysis for the stiffness constant ${{K}_{\text{A}}}(T)$ , defined in equation (151) in exactly the same way, using a Debye–Hückel [78] approximation and the dislocation Hamiltonian of equation (166) to obtain

Equation (174)

This means that the orientational stiffness constant ${{K}_{\text{A}}}(T)=\infty $ for an adsorbate on an incommensurate substrate, which, in turn, means that there is true long range orientational order for all $T>{{T}_{\text{m}}}$ and that there is no high temperature transition to the expected isotropic fluid. Thus, the experimental investigation seems to be a very difficult proposition and there is an excellent discussion in the review article by Strandburg [142] with extensive references to many aspects of the melting problem.

3.3. Experimental verification of the KTHNY melting scenario

This difficulty was finally overcome in 1999 [157] by trapping a monolayer of colloidal particles at an air/water interface, doped with paramagnetic ions in a magnetic field B normal to the interface. The interaction between the particles is characterized by $\Gamma=\left({{\mu}_{0}}/4\pi \right)\left(\chi {{B}^{2}}\right){{(\pi n)}^{3/2}}/\left({{k}_{\text{B}}}T\right)$ where n is the areal number density and χ the magnetic susceptibility. This system is as close to the conditions of the theory as it is possible to get as the substrate is completely isotropic and free of any destructive perturbations. The particle positions are recorded by video camera so that all correlation functions can be readily calculated and compared with theory. The outcome is that, in this system, measurements agree quantitatively with the KTHNY theory in all respects with no adjustable parameters [155, 156].

One of the major experimental difficulties is equilibrating the system, despite it being only about 105 particles. In the high field crystalline phase, the equilibration time is several days [157] even with the system being jiggled by applying a small ac ${{B}_{||}}$ in the plane of the system to anneal out the large out of equilibrium concentration of defects such as grain boundaries, interstitials and vacancies. Eventually, the system consists of a single domain with less than 0.1% of dislocations. To melt this 2D crystal, the temperature ${{\Gamma}^{-1}}$ is increased by decreasing B in small steps equilibrating for about one hour after each step [157]. This procedure results in ${{C}_{\text{G}}}\left(\mathbf{r}\right)$ decaying as a power law for $T\leqslant {{T}_{\text{m}}}$ and exponentially for $T>{{T}_{\text{m}}}$ , and similarly for ${{C}_{6}}\left(\mathbf{r}\right)$ , as shown in figure 20

Equation (175)

where ${{\eta}_{\mathbf{G}}}(T)$ is given in (143), ${{\xi}_{+}}(T)$ in (146), ${{\eta}_{6}}(T)$ in (157) and ${{\xi}_{6}}(T)$ in (160). The experimental results for ${{C}_{\mathbf{G}}}\left(\mathbf{r}\right)$ and ${{C}_{6}}\left(\mathbf{r}\right)$ are shown in figure 20 and are consistent with the theoretical predictions. Note that ${{C}_{\mathbf{G}}}\left(\mathbf{r}\right)$ in the solid phase decays as a power law ${{r}^{-{{\eta}_{\mathbf{G}}}(T)}}$ with ${{\eta}_{\mathbf{G}}}\left({{T}_{\text{m}}}\right)=1/3$ . This agrees with theory because the measured power law is actually ${{r}^{-{{\eta}_{{{\text{G}}_{0}}}}}}$ where G0 is the magnitude of the smallest reciprocal lattice vector for a hexagonal lattice ${{G}_{0}}=4\pi /\left({{a}_{0}}\sqrt{3}\right)$ .

Figure 20.

Figure 20. Correlation functions ${{C}_{\text{G}}}\left(\mathbf{r}\right)$ (left) and ${{C}_{6}}\left(\mathbf{r}\right)$ (right). The solid lines have slopes 1/3 (left) and 1/4 (right). At $T=T_{\text{m}}^{-}$ , data for ${{C}_{\text{G}}}\left(\mathbf{r}\right)\sim {{r}^{-1/3}}$ and, at $T=T_{I}^{-}$ , ${{C}_{6}}\left(\mathbf{r}\right)\sim {{r}^{-1/4}}$ . Reprinted from [157] with permission. Copyright 1999 American Physical Society.

Standard image High-resolution image

Measured values of the Young's modulus ${{\tilde{K}}_{\text{R}}}(T)$ compared with theory are shown in the upper panel of figure 21 and for the elastic constants ${{\tilde{\mu}}_{\text{R}}}(T)$ and ${{\tilde{\lambda}}_{\text{R}}}(T)+2{{\tilde{\mu}}_{\text{R}}}(T)$ in the lower panel which also agree well with theory. The experimental systems are electrons on 4He [150] and paramagnetic colloidal particles at an air/water interface [144] as these are the closest to the conditions of theory and are free from unwanted substrate effects. As shown in (145), for an incompressible lattice, ${{\eta}_{{{\text{G}}_{0}}}}=1/3$ which is close to the theoretical value for particles interacting by a 1/r3 repulsive potential. As shown in [144], the exponent

Equation (176)

so that, at $T_{\text{m}}^{-}$ , ${{\Gamma}_{m}}{{\eta}_{{{\text{G}}_{0}}}}\left(T_{\text{m}}^{-}\right)=1.496$ , as shown in figure 22.

Figure 21.

Figure 21. Upper: Young's modulus ${{K}_{\text{R}}}(T)$ . Lower: Elastic constants, theory and experiment. Reprinted from [143, 144] with permission. Copyright 2005 IOP Publishing and copyright 2004 American Physical Society.

Standard image High-resolution image
Figure 22.

Figure 22. The exponent ${{\eta}_{{{\text{G}}_{0}}}}(T)$ from the measured elastic constants ${{\mu}_{\text{R}}}(T)$ and ${{\lambda}_{\text{R}}}(T)$ . The theoretical prediction (solid line) compared with experiment and Monte Carlo for $\eta \gamma $ compared with experiment and simulation. Reprinted from [144] with permission. Copyright 2005 IOP Publishing.

Standard image High-resolution image

One of the main interests in melting in 2D is the order of the transition. It is generally agreed that the solid/liquid transition in 3D is first order characterized by a latent heat. In fact, the melting transition in 3D is considered as the paradigm of a discontinuous first order transition. The KTHNY theory of melting in 2D predicting successive continuous transitions from the low T solid to the high T isotropic liquid came as a heretical claim which needed to be demonstrated by experiment to be false. However, it turned out to be very difficult to either verify or falsify the KTHNY theory because there were no experimental systems available which obeyed the very stringent conditions of the theory which requires the adsorbate layer to be constrained by a featureless substrate potential of the form $V(x,y,z)=-{{V}_{0}}\delta \left(z-{{z}_{0}}\right)$ with ${{V}_{0}}\gg {{k}_{\text{B}}}T$ which restricts the adsorbate to a flat layer at z  =  z0 where the substrate surface is at z  =  0. The theory requires that the potential V(x, y, z) is independent of the substrate coordinates (x, y) and this can be realized only in two situations: (i) electrons trapped in a layer above the surface of 4He [150, 153, 154] and (ii) paramagnetic colloidal particles trapped at a water/air interface [157].

Experiments on orientational order in the hexatic phase have also been performed on a layer of paramagnetic colloidal particles interacting by a 1/r3 repulsive energy [155, 157, 158] with very good agreement with the theoretical predictions [129]. Excellent reviews summarizing the present status of the transitions between the three phases of particles in two dimensions with 1/r and 1/r3 repulsive interactions are in [144, 156]. Unfortunately, orientational order and the hexatic phase cannot be studied in the electrons on helium system because measurements can be made only on the frequency spectrum of resonances of the coupled electron—ripplon modes [95, 150, 151]. The comparison between theory and experiment for the orientational stiffness ${{\tilde{K}}_{\text{A}}}(T)$ is shown in figure 24. The theoretical renormalization group predictions of the KTHNY theory are the solid line and the experimental measurements are the points. To the author's eye, the agreement is very good, considering that there are no adjustable parameters in the fits.

The orientational correlation function ${{G}_{6}}\left(\mathbf{r}\right)$ is investigated and the results are summarized in figure 23 where it is seen that ${{G}_{6}}(r)\sim {{r}^{-{{\eta}_{6}}(T)}}$ for $r/a\leqslant 30$ , which is limited by the sample size. ${{G}_{6}}\left(\mathbf{r}\right)$ is consistent with the theoretical predictions with a power law decay with ${{\eta}_{6}}(T)\leqslant 1/4$ for ${{T}_{\text{m}}}<T\leqslant {{T}_{I}}$ and exponential decay for T  >  TI. The correlation length ${{\xi}_{6}}(T)$ is shown in figure 24(a) and is consistent with the theoretical prediction ${{\xi}_{6}}(T)\sim {{\text{e}}^{b{{t}^{-1/2}}}}$ although the experimental points are scattered about the theoretical solid line. The orientational stiffness constant, ${{\tilde{K}}_{\text{A}}}(T)$ , is shown in figure 24(b) where agreement with theory, $\tilde{K}_{\text{A}}^{\text{R}}\left(T_{I}^{+}\right)=72/\pi $ and $\tilde{K}_{\text{A}}^{\text{R}}\left(T_{m}^{+}\right)\sim \text{exp}\left(b{{t}^{-1/2}}\right)$ .

Figure 23.

Figure 23. Orientational correlation function ${{G}_{6}}\left(\mathbf{r}\right)$ . The top three curves have ${{G}_{6}}\left(\mathbf{r}\right)=\text{constant}$ , the next two ${{G}_{6}}\left(\mathbf{r}\right)\sim {{r}^{-{{\eta}_{6}}(T)}}$ with ${{\eta}_{6}}(T)\leqslant 1/4$ and the bottom two ${{G}_{6}}\left(\mathbf{r}\right)\sim {{\text{e}}^{-r/{{\xi}_{6}}(T)}}$ . Reprinted from [158] with permission. Copyright 2007 American Physical Society.

Standard image High-resolution image
Figure 24.

Figure 24. (a) Orientational correlation length ${{\xi}_{6}}(T)$ . The solid line is from theory ${{\xi}_{6}}(T)\sim {{\text{e}}^{b{{t}^{-1/2}}}}$ . (b) Orientational stiffness constant ${{\tilde{K}}_{\text{A}}}(T)=\beta {{F}_{\text{A}}}$ compared with theory. Reprinted from [158] with permission. Copyright 2007 American Physical Society.

Standard image High-resolution image

On balance, the agreement with theory verifies the KTHNY theory of melting at least for this particular system as there are no contradictions. However, it must be mentioned that the actual critical region in which the exponential divergence of the correlation length ${{\xi}_{+}}(t)\sim {{\text{e}}^{b{{t}^{-\tilde{\nu}}}}}$ sets in is estimated to be $\mathcal{O}\left({{10}^{8}}\right)$ lattice spacings [164]. This is much larger than accessible experimental sizes which casts grave doubt about the significance of the fitting of data to the theoretical form of ${{\xi}_{+}}$ ! In particular, because of the limited sample sizes, these experiments are unable to exclude weak first order transitions seen in large scale simulations of hard disks [159161].

3.4. First order transitions in planar rotor and melting models

The transition the 2D planar rotor and melting models are argued to be driven by vortex and dislocation unbinding [19] which lead to continuous transitions. However, symmetry arguments do not exclude the possibility that, in some systems, these defect driven transitions are first order. The earliest study of melting in 2D which gave convincing evidence of a transition was carried out by Alder and Wainwright [31] and there have been many others since then.

However, a recent very large scale simulation of a hard disk system with N  =  10242 disks [160, 161] obtains results which agree with the theoretical predictions for the solid—hexatic transition but show that the hexatic—isotropic fluid transition is weakly first order with a discontinuity in the density of about 0.015. Although this scenario sounds unlikely, it is entirely possible. It is known that the KT transition can be made first order by reducing the vortex core energy ${{E}_{\text{c}}}$ to below some critical value [165167]. In [165], a modified planar rotor Hamiltonian on a square lattice was proposed

Equation (177)

This is precisely the planar rotor model when p  =  1 which has a continuous transition driven by vortex unbinding while, for large p, $V(\theta )\approx 2J$ for $\theta \geqslant \pi /p$ and $V(\theta )\approx \frac{1}{2}J{{p}^{2}}{{\theta}^{2}}$ for $\theta \ll \pi /p$ [166]. The physical argument for a first order transition for $p\gg 1$ is that the Hamiltonian of (177) resembles that of a Potts model [105] with a large number of states which is known to undergo a first order transition [107]. Vortex unbinding is at ${{k}_{\text{B}}}{{T}_{\text{KT}}}\approx J{{p}^{2}}$ and vacancy condensation at ${{k}_{\text{B}}}{{T}_{\text{D}}}\approx V(\pi )=2J$ . For p sufficiently large, one expects that ${{T}_{\text{D}}}<{{T}_{\text{KT}}}$ so that the continuous KT transition is preempted by a first order vacancy condensation [166]. These physical arguments are verified by simulations [165167] although the system sizes are somewhat limited and proved rigorously in [168]. Also, the vortex core energy ${{E}_{\text{c}}}$ is reduced as p increases which indicates that a defect description with an appropriate defect fugacity y0 is sufficient to account for both types of transition. In particular, this can explain the first order hexatic-isotropic fluid transition observed in simulations of the hard disk system [159161]. It is worth noting that this simulation of up to N  =  220 hard discs is, in fact, larger than any 2D experimental system. Of course, it differs somewhat from a real particle system since these have interactions beyond the hard core repulsion.

In some of the earliest Monte Carlo (MC) simulations of hard spheres and discs interacting by a Lennard-Jones potential $V(r)=4\epsilon \left[{{(\sigma /r)}^{12}}-{{(\sigma /r)}^{6}}\right]$ [26], it was found that a system of N  =  56 particles in 2D and up to 256 in 3D with periodic boundary conditions showed indications of a gas/liquid transition in 3D but not in 2D. Studies of the 3D hard sphere system by MC [27] and by molecular dynamics (MD) [2830] agreed on the equation of state and on a first order transition in 3D with small finite size effects. Later, a MD simulation of N  =  870 hard disks in 2D was done [31] where a first order melting/freezing transition was observed. Since these early studies, there has been over fifty years of numerical study on the apparently simple hard disk system culminating in the massive simulations by Krauth and coworkers [159161]. It was discovered that 2D melting is neither first order nor KTHNY but has the low T crystal melts to a hexatic fluid by a continuous KTHNY transition followed by a very weak first order transition between the hexatic and isotropic fluids.

Since Alder and Wainwright's simulation of the hard disk system [31], many simulations of 2D hard disk and Lennard-Jones systems have been performed, and excellent reviews of the situation before 1985 are by [142, 169]. The great majority of simulations concluded that melting in 2D is by a single first order transition, with few exceptions e.g. [170] who pointed out that his simulations were unable to distinguish between a single weak first order and the KTHNY scenario. It is of interest to note that the existence of the hexatic fluid phase has been confirmed by simulation only very recently [159, 161]. Even relatively large simulations have not been able to yield an unambiguous confirmation of the existence of a hexatic phase [171178] and conclude that 2D melting is a single first order transition with the exception of [176] where the conclusion was that 2D melting takes place by a single continuous transition. This is mainly due to size limitations, the largest system simulated was N  =  214 disks, but also the underlying expectation that 2D melting should be similar to 3D first order melting. Thus, usually only orientational order was followed and translational order assumed to be slaved to this but, even when the two orders were considered, the hexatic phase was not seen unambiguously. In a simulation of 220 hard disks, Jaster [179] obtained evidence of the hexatic phase by finite size scaling but the nature of the hexatic—isotropic fluid transition could not be determined [180]. This was finally settled very recently by Krauth and coworkers [159, 161], who found a very narrow hexatic phase reached from the solid by KTHNY melting followed by a very weak first order transition to the isotropic fluid. The first order nature is verified by the finite size scaling of the interface free energy $\Delta f(N)\sim {{N}^{-1/2}}$ for particle number N  =  2n, n  =  14, 16, 18, 20, provided the correlation length $\xi \ll \sqrt{N}$ [181]. The hard disk simulations [160, 161] obey this criterion thus establishing the first order nature of the hexatic-isotropic fluid transition. The crystal-hexatic transition was also checked to be KTHNY by the behavior of the density–density correlation function, $\langle {{\rho}_{{{\text{G}}_{0}}}}\mathbf{r}{{\rho}_{{{\text{G}}_{0}}}}(0)\rangle \sim {{r}^{-\eta}}$ in the solid phase and $\sim {{\text{e}}^{-r/\xi}}$ in the hexatic phase with $\eta =1/3$ at the hexatic-crystal transition [160, 161].

An elastic solid can be described by interacting dislocations and disclinations, provided the concentration of point defects such as vacancies and interstitials is not too large [19, 128, 129]. These point defects are irrelevant in the renormalization group sense and can be absorbed into the elastic constants $\mu,\lambda $ . Thus, a natural way to study melting numerically is to simulate the dislocation free energy of (121) with parameters K0 and dislocation core energy ${{E}_{\text{c}}}$ . When ${{E}_{\text{c}}}\to 0$ , dislocations are expected to form grain boundaries between crystallites of different orientation and Chui [182, 183] constructed a theory of grain boundary melting which gives a first order transition where both translational and orientational order vanish simultaneously. Saito [184, 185] simulated a system of dislocations of (121) for different values of the core energy ${{E}_{\text{c}}}$ . It is found that, for large ${{E}_{\text{c}}}>E_{\text{c}}^{\ast}\approx 1.22{{K}_{0}}$ a continuous KYHNY melting transition occurs at ${{K}_{\text{R}}}=16\pi $ [129] but when ${{E}_{\text{c}}}<E_{\text{c}}^{\ast}$ , there is first order melting directly to an isotropic fluid [184, 185]. When ${{E}_{\text{c}}}<E_{\text{c}}^{\ast}$ , the dislocations organize themselves into grain boundary loops and first order melting results [182, 183]. Of course, these point defects are important for the dynamics of the collective defects like dislocations. Dislocation glide conserves particle number but climb or motion parallel to the Burgers vector requires diffusion of vacancies and interstitials and is therefore very slow [186]. Our purely elastic description does not take such effects into consideration and to do this would need more of a microscopic description.

In the simpler system of superfluid 4He with some non-superfluid defects, 3He atoms, there is a continuous transition for low impurity concentration which becomes first order for larger concentrations [86]. In this work, the impurities are incorporated into the planar rotor model on a lattice by introducing a ti  =  0, 1 to represent a 3He atom by ti  =  0 and a 4He atom by ti  =  1. The resulting Hamiltonian is [86]

This was studied on a triangular lattice by an approximate Migdal [187] type bond moving renormalization group procedure due to Kadanoff [188]. As expected, both continuous and discontinuous transitions appeared. Although never carried out with impurities, a calculation along these lines has been performed by Nelson [127]. It seems that this would be a relatively simple way to study the effects of point defects (vacancies and interstitials) on the continuous dislocation and disclination unbinding transitions and to explore possible melting scenarios. This goes beyond the purely elastic theories of 2D melting which require the introduction of disclinations to induce a first order transition. One outstanding problem with this approach is that the first order transition induced by disclinations [189, 190] requires going beyond linear elasticity and causes the simultaneous destruction of both translational and orientational order. This seems rather restrictive and certainly does not always happen in real systems. A typical phase diagram from purely elastic theory is sketched in figure 25.

Figure 25.

Figure 25. Reprinted with permission from [193]. Copyright 1988 Elsevier. A schematic phase diagram for melting in 2D in a purely elastic theory. The continuous solid/hexatic and hexatic/isotropic fluid transitions are labelled KT. The dashed line marks the continuation of the solid/hexatic transition to ${{l}^{2}}<l_{\text{c}}^{2}$ where the sold line is a first order solid/isotropic fluid transition. See also [195, 196]

Standard image High-resolution image

In a series of papers [189196], Kleinert and coworkers developed and simulated a model for 2D melting including an important additional parameter which allows for first order melting within a theory based on elasticity. By allowing for higher order elastic effects [195] one can incorporate local rotations

Equation (178)

where uij is the strain tensor and $\omega =\frac{1}{2}\left({{\partial}_{1}}{{u}_{2}}-{{\partial}_{2}}{{u}_{1}}\right)$ is the local rotation field. The parameters l2 and ${{l}^{\prime 2}}$ appear in $\omega \left(\mathbf{q}\right)$ relations as [195]

Equation (179)

However, Kleinert [195, 196] has shown that the parameter ${{l}^{\prime 2}}$ is irrelevant for melting and considers only l2, which is a measure of rotational stiffness. A model combining both mechanisms leading to first order melting [192] and [86] might be more realistic than existing models.

Most simulations of melting of an elastic solid have been performed in the dual roughening representation, which is an exact transformation of the interacting dislocations and disclinations with long range interactions into a roughening model on the dual lattice with short range interactions [142, 193, 197, 198]. These models have a phase diagram of figure 25 with both continuous melting or a first order transition directly to the isotropic fluid and have a definite computational advantage of nearest neighbor interactions but incorporating local defects is more awkward.

The KTHNY melting theory inspired a large experimental effort to verify or falsify the two continuous transition scenario, summarized by Strandburg in a comprehensive review [142]. There have been several studies of liquid crystal films [96103, 121, 199201]. In some of these, melting seems to proceed according to the KTHNY scenario [121] but these freely suspended systems are much richer than theory suggests with many different possible phases depending on temperature and film thickness [201]. One interesting observation is that, as the film reaches its minimum thickness of two layers, a hexatic phase is observed which seems to behave as predicted by theory [123, 202], although the evidence for a separate hexatic phase is not conclusive.

Adsorbed gases on substrates provide a huge variety of 2D systems but the interplay of the periodic elastic adsorbed layer with reciprocal lattice vectors $\mathbf{G}$ and the periodic substrate with reciprocal lattice vectors $\mathbf{M}$ complicates things dramatically. The adsorbed over layer can be commensurate or incommensurate with the substrate, depending on the relationship of $\mathbf{G}$ and $\mathbf{M}$ [128, 129]. An early but comprehensive theoretical review on commensurate/incommensurate transitions is [163] and one on experiments on rare gases on graphite [203]. Graphite is one of the most widely used substrates because it can provide fairly large atomically flat surfaces of linear size  ∼1600 Å. Favorite adsorbates are rare gases such as Ar, Kr, Ne and Xe which are believed to interact by a simple Lennard-Jones interaction and differ only in size and monolayers of these have been extensively studied by diffraction of synchrotron radiation [204206]. Xenon adsorbed on graphite has been extensively studied [207] where both first order melting of the incommensurate solid at lower coverage and continuous melting at higher coverage was seen. Krypton on graphite has been studied extensively [206, 208214] and a tongue of fluid found separating the commensurate and incommensurate phases as shown in figure 26. High resolution scattering experiments are consistent with a smooth continuous F-IC transition [208] but the resolution is insufficient to make detailed fits to theory [215]. However, the scattering data seems to be inconsistent with KTHNY melting [208].

Figure 26.

Figure 26. Phase diagram of Krypton adsorbed on graphite showing the fluid (F), the $\sqrt{3}\times \sqrt{3}$ commensurate phase (C), the reentrant fluid (RF) and the incommensurate solid (IC). The IC/RF transition is the relevant melting transition here. Reproduced from [208] with permission. Copyright 1984 American Physical Society.

Standard image High-resolution image

Argon is very similar to Krypton except it has a slightly smaller atomic size and always forms an incommensurate solid overlayer on graphite with a phase diagram of figure 27 left. Diffraction studies of the melting transition around monolayer coverage are consistent with KTHNY melting but the correlation length is equally well fitted by a conventional power law $\xi (t)={{\xi}_{0}}{{t}^{-\nu}}$ [216]. It must be remembered that length scales are limited to $\xi =\mathcal{O}\left({{10}^{3}}\right)$ Å. These results agree with an early specific heat measurement [217] but disagree with a later specific heat measurement [218] and earlier diffraction experiments [204]. A later investigation [219] is also consistent with KTHNY melting and also with the prediction that the solid over layer is rotated from the substrate axes [220, 221] but again detailed comparison with theory is difficult because of insufficient experimental resolution. A study using high resolution vapor pressure isotherms [222] yields evidence in favor of continuous melting by a two step process which is constant with KTHNY melting. Although not conclusive, this study does resolve some discrepancies between the older measurements, which makes the KTHNY scenario more likely for monolayer Argon on graphite.

Figure 27.

Figure 27. Phase diagrams of argon (left) and xenon (right) adsorbed on graphite. Reproduced from [216] (argon) and [224] (xenon) with permission. Copyright 1987 American Physical Society, and copyright 1983 Elsevier.

Standard image High-resolution image

Xenon adsorbed on graphite has a phase diagram of figure 27 right and has also been studied by diffraction [207, 219, 223225] and by thermodynamic measurements [226228]. The synchrotron diffraction experiments [207, 223, 224] are consistent with first order melting for low coverages which crosses over to KTHNY continuous melting at higher coverages. Thermodynamic measurements of the compressibility of the overlayer [227] are consistent with this. However, a high precision heat capacity and compressibility measurements [228] are not consistent with a crossover to continuous KTHNY melting and claim that melting is always first order. The scattering experiment of Rosenbaum et al [225] showed that melting at high coverages is well described by KTHNY melting in the presence of a weak ${{h}_{6}}\text{cos}6\theta $ substrate potential. A study of melting of Xenon adsorbed on Ag(1 1 1) [229] agrees with the KTHNY scenario. In fact, the 6-fold substrate potential of the silver substrate is extremely small [229] and this model fits the scattering data very well. The observed hexatic order of the fluid seems entirely due to the Xenon overlayer. The only effect of the substrate is to align the 6-fold hexatic axes by steps in the substrate [229] and the hexatic orientational order is intrinsic to the Xenon over layer. The remaining rare gas, Neon, on graphite has also been studied [230232] but in much less detail than the others. In view of the recent simulations of the hard disk system [159] it seems that the conclusions from experiments must be viewed with some caution because of the very limited system sizes and the difficulty of distinguishing continuous from weak first order transitions.

Many other adsorbates on graphite have been studied by thermodynamic measurements with many complicated phase diagrams proposed. A neutron scattering study of melting of ethylene adsorbed on graphite has yielded evidence of a continuous transition [233] and compared to methane on the same substrate. Both films melt from incommensurate solids which rules out substrate heterogeneity being responsible for the apparent continuity in ethylene. Methane films show no signs of this broadening. Thus, the broadening of the transition is taken to be intrinsic to the overlayer melting by successive continuous transitions but the resolution does not permit verification that this is KTHNY melting [233].

3.5. Melting of arrays of anisotropic particles

Another relevant system in 2D is a freely suspended film of liquid crystal [96, 100, 132] on which measurements are possible. There are a number of theoretical papers generalizing earlier work on melting of arrays of point particles to 2D arrays of anisotropic particles [123, 202, 234]. Ostlund and Halperin [202] discuss dislocation mediated melting of 2D arrays of rods tilted from the normal to the layer. Note that, if the tilt angle is zero, the system can be described by an array of point particles so the theory is the same as the previous section. However, for tilted rods as in figure 28(a), the projection of the rods into the plane is a set of directed arrows arranged in a distorted triangular lattice. Two obvious physical cases considered are (i) the projection of the molecular axes tend to point along one of the six nearest neighbor bond directions as sketched in figure 28(b) and (ii) tend to point intermediate between two bonds as in figure 28(c). A very important consequence of the molecular tilt is that, even if in the absence of tilt, the molecules would form a regular hexagonal lattice, the molecules will form a uniaxial hexagonal lattice [123].

Figure 28.

Figure 28. (a) Tilted molecules in a Smectic C monolayer. (b) Top view of monolayer of tilted molecules where arrows are upper ends. The molecular axes tend to point along one of the nearest neighbor directions so that arrows tend to form chains. The distorted hexagonal lattice formed by the centers of the molecules is shown by dotted lines. (c) Same as (b) except that the projections of the molecular axes lie halfway between nearest neighbor bonds. Reprinted from [202], with permission. Copyright 1981 Elsevier.

Standard image High-resolution image

To study the dislocation mediated melting of a 2D solid composed of rodlike molecules tilted at angle θ from the plane and the subsequent transitions which is relevant for freely suspended films of smectic liquid crystals [96, 100, 121, 131, 132], we must go through the same steps as for the melting of a hexagonal lattice. Unfortunately, even the first step is rather complicated because the energy of a set of dislocations in a uniaxial hexagonal lattice was not known in the late 1970's. The most explicit calculation is in [235] but, as pointed out by Ostlund and Halperin [202] the result depended on the choice of a cutoff and must be incorrect. In appendix A of [202] the dislocation energy is calculated without a cutoff and is more likely to be correct. In this review, I will try to focus on the salient points and leave all the gory technical details to the references [202].

The first complication is to compute the energy of a set of dislocations embedded in the elastic medium with the uniaxial symmetry of the underlying distorted hexagonal lattice shown in figure 29. In such a system, there are four different nonzero elastic constants or, equivalently, four distinct finite compliances. The compliance tensor Sijkl is the inverse of the elastic tensor Cijkl or ${{C}_{ijkl}}{{S}_{klmn}}=\frac{1}{2}\left({{\delta}_{im}}{{\delta}_{jn}}+{{\delta}_{in}}{{\delta}_{jm}}\right)$ . The finite compliances are ${{S}_{1111}},{{S}_{2222}},{{S}_{1122}},{{S}_{1212}}$ where the axes are shown in figure 29.

Equation (180)

where $\alpha,\beta $ denote one of the six dislocations of figure 29 in an anisotropic hexagonal lattice, i, j  =  1, 2 are the two independent directions in 2D and ${{\mathbf{r}}^{\alpha}}$ is the position of the dislocation of type α. There are two independent Burger's vectors ${{\mathbf{b}}_{\text{I}}}$ and ${{\mathbf{b}}_{\text{II}}}$ . The interaction energy of a pair of  ±  type α dislocations separated by $\mathbf{r}$ is ${{K}_{\alpha}}=b_{i}^{\alpha}b_{j}^{\alpha}{{E}_{ij}}\left(\mathbf{r}\right)$ , where ${{E}_{ij}}\left(\mathbf{r}\right)={{K}_{ij}}\text{ln}(r/a)+{{V}_{ij}}(\theta )$ is the interaction energy which is a very, complicated expression given in appendix A of [202].

Figure 29.

Figure 29. Left: (a) Direct lattice vectors and elementary Burgers vectors for the lattice of figure 28(b) and (b) reciprocal lattice vectors. Right: the same but for figure 28(c). Reprinted from [202], with permission. Copyright 1981 Elsevier.

Standard image High-resolution image

The solid phase has dislocations which are tightly bound in pairs and when the temperature is raised, the pairs unbind and the crystalline order is destroyed. This melting occurs when $\Delta F=\Delta E-T\Delta S$ where $\Delta S={{k}_{\text{B}}}\text{ln}\left({{L}^{2}}/{{a}^{2}}\right)$ and $\Delta E=\pi K\text{ln}(L/a)$ where K is determined by the lattice spacing and elastic constants, L is the size of the system and a the lattice constant [19]. The melting temperature is given by $\Delta F=0$ or $\pi K=2{{k}_{\text{B}}}T$ . There are two types of dislocation and two values ${{K}_{\text{I}}}$ and ${{K}_{\text{II}}}$ in an anisotropic solid, as shown in figure 29. Thus, melting of the solid will be controlled by the smaller of ${{K}_{\text{I}}},\,{{K}_{\text{II}}}$ which is, in turn, controlled by the shorter of the elementary Burger's vectors ${{\mathbf{b}}_{\text{I}}}$ and ${{\mathbf{b}}_{\text{II}}}$ . Following the discussion of [123, 202, 234], one can study the melting of the anisotropic solid by the renormalization group procedure of section 3.1. Following [202], the compliance tensor Sijkl is the inverse of the elastic tensor Cijkl so that ${{C}_{ijkl}}{{S}_{klmn}}=\frac{1}{2}\left({{\delta}_{im}}{{\delta}_{jn}}+{{\delta}_{in}}{{\delta}_{jm}}\right)$ and, to $\mathcal{O}\left(\,{{y}^{2}}\right)$ , and Sijkl is given by

where the interaction energy of a  ±  pair of type α dislocations separated by $\mathbf{r}$ is

The RG equations are derived from this by the standard method of rescaling $a\to a{{\text{e}}^{\delta l}}$ to obtain [202]

Equation (181)

An expression for ${{V}_{ij}}(\theta )$ is given in appendix A of [202]. The quantities $F_{\alpha}^{ij}$ and ${{A}_{\alpha \beta}}$ depend on the compliances which have finite fixed point values and differ from these by $\mathcal{O}\left(\,{{y}^{2}}\right)$ . Thus, one can consider these as l-independent constants [202].

For type I melting, ${{K}_{\text{II}}}>4$ and ${{K}_{\text{I}}}\approx 4$ so that the fugacity ${{y}_{\text{II}}}\to 0$ because $\left(\frac{{{K}_{\text{II}}}}{2}-2\right)>0$ and the flow equations involving ${{y}_{\text{I}}}$ become

Equation (182)

Near the melting temperature, the compliances S1122 and S2222 are dominated by analytic terms since their RG flows are not controlled by the fugacity ${{y}_{\text{I}}}$ . The compliances ${{S}_{1111}},{{S}_{1212}}$ are expanded as

Equation (183)

Here, $S_{1111}^{\ast}$ is the fixed point value of S1111 at melting and is equal to its renormalized value at ${{T}_{\text{m}}}$ . D(l) is the deviation from the incoming separatrix with $D(0)=\,|t|$ and ${{y}_{\text{I}}}$ is the value of the fugacity on the incoming separatrix.

Below melting, $T\leqslant {{T}_{\text{m}}}$ , after some further algebra [202], one can derive the flow equations for ${{y}_{\text{I}}}(l)$ and D(l)

Equation (184)

where the slope m  <  0 [202]. For $|t|\ll 1$ , ${{y}_{\text{I}}}(l)\sim {{l}^{-1}}$ for $l\gg 0$ . Also $D(l){{y}_{\text{I}}}(l)=D(0){{y}_{\text{I}}}(0)$ which allows one to define a value l* such that $D\left({{l}^{\ast}}\right)={{y}_{\text{I}}}\left({{l}^{\ast}}\right)$ such that for l  >  l*, the fugacity ${{y}_{\text{I}}}(l)\to 0$ so that one can write $|t|\propto D(0)\propto {{\left({{l}^{\ast}}\right)}^{2}}$ . In turn, this observation, allows one to define a length scale ${{\xi}_{-}}(t)$ and the singularities in the compliances

Equation (185)

where b  >  0 and ${{b}^{\prime}}>0$ are nonuniversal constants $\mathcal{O}(1)$ . Thus for $T\leqslant {{T}_{\text{m}}}$ , the compliances S1111 and S1212 have $\sqrt{|t|}$ singularities while S1122 and S2222 have undetectable essential singularities $\xi _{S}^{-2p}$ at ${{T}_{\text{m}}}$ , since ${{K}_{\text{II}}}>4$ [202].

Above melting, $T>{{T}_{\text{m}}}$ , the RG trajectories for ${{y}_{\text{I}}}(l)$ follow the incoming separatrix up to $l=l_{1}^{\ast}$ , break away and join the outgoing separatrix ${{y}_{\text{I}}}(l)={{m}_{+}}\left({{S}_{1111}}(l)-S_{1111}^{\ast}\right)$ for another $l_{2}^{\ast}$ RG iterations. ${{y}_{\text{I}}}\left(l_{2}^{\ast}\right)=\mathcal{O}(1)$ but is still small. Thus, we see that the length scale ${{\xi}_{S}}\sim \text{exp}\left(l_{1}^{\ast}+l_{2}^{\ast}\right)$ is the typical separation of free dislocations of type I. The fugacity ${{y}_{\text{II}}}(l)$ is decreasing exponentially for $l\leqslant l_{1}^{\ast}+l_{2}^{\ast}$ . Thus, we can integrate out the type I dislocations by a Debye–Hückel approximation to obtain an effective Hamiltonian for the type II dislocations at scale $l_{S}^{\ast}=l_{1}^{\ast}+l_{2}^{\ast}$ , which is the Hamiltonian for a 2D smectic liquid crystal [202, 234]

Equation (186)

At first sight, it seems that at this length scale one can treat both type I and type II dislocations by a Debye–Hückel approximation but ${{y}_{\text{II}}}\left(l_{S}^{\ast}\right)\ll 1$ so that the discrete nature of ${{\mathbf{b}}_{\text{II}}}$ must still be accounted for [234]. Thus, a further $l_{3}^{\ast}$ iterations are performed so that

Equation (187)

At this length scale, the system becomes a gas of free dislocations of all types and one can use a Debye–Hückel approximation. At this scale, the Frank constants

Equation (188)

At length scale L with $L<{{\xi}_{S}}(t)$ , all dislocations are bound and the system responds like a solid while, if ${{\xi}_{S}}<L<{{\xi}_{N}}$ , it has a smectic like response. At still larger scales, ${{\xi}_{N}}<L<{{\xi}_{I}}$ , it behaves like a nematic with two Frank constants ${{K}_{y}}/{{K}_{x}}\propto \xi _{S}^{2p}$ . For still larger scales, higher order terms like ${{\left(\nabla \centerdot \mathbf{n}\right)}^{4}}$ in the Hamiltonian will renormalize ${{K}_{x}},{{K}_{y}}$ to equality for $L>{{\xi}_{I}}\sim \text{exp}\left(c\xi _{S}^{2}\right)$ [236]. This melting scenario is called type I melting [202] with a phase diagram sketched in figure 27, left. When ${{K}_{\text{II}}}>{{K}_{\text{I}}}$ , the smectic-like phase is absent and the phase diagram is sketched in figure 27, right.

3.6. Superconductivity in two dimensions

Another system which, at first sight, seems that it should be a good example of experimentally accessible Kosterlitz–Thouless physics is a thin film of superconductor. In this system, it is clear that the physics will be controlled by vortices in the phase of the superconducting order parameter $\Psi\left(\mathbf{r}\right)=\,|\Psi\left(\mathbf{r}\right)|{{\text{e}}^{\text{i}\phi \left(\mathbf{r}\right)}}$ where ϕ is the phase of the condensate wavefunction. This is exactly analogous to the superfluid 4He situation. The main difference is that the superconducting condensate is charged which implies that the all important vortex–vortex interaction is screened from the $\text{ln}r$ form of the neutral superfluid to 1/r due to screening currents flowing round the vortices [237, 238].

This is best discussed from the Ginzburg–Landau form of the free energy for a charged condensate of Cooper pairs [239]

Equation (189)

$\Psi\left(\mathbf{r}\right)$ is the coarse grained Cooper pair wave function which is non zero only in the plane of the film z  =  0, $\mathbf{A}\left(\mathbf{r},z\right)$ is the vector potential which is finite outside the plane of the film and $\mathbf{B}\left(\mathbf{r},z\right)=\nabla \times \mathbf{A}$ . The parameter r(T)  <  0 for T  <  T0 the BCS temperature for the onset of the formation of Cooper pairs and u is positive and T independent. Here $\mathbf{r}=(x,y)$ is a point in the film in the z  =  0 plane, the Cooper pair mass ${{m}^{\ast}}=2{{m}_{e}}$ and e*  =  2e. The essential difference between the neutral 4He superfluid and the charged superfluid of a superconductor is that a vortex in a superconductor has a circulating electric current producing an associated magnetic field $\mathbf{B}\left(\mathbf{r},z\right)$ which extends outside the film at $\left(\mathbf{r},0\right)$ .

The behavior of a superconducting film was first discussed by Pearl [237] who found that the circulating current of a vortex at $\mathbf{r}=0$ falls off as 1/r2 instead of exponentially as in a bulk superconductor. Also, the screening length in a film is ${{\lambda}_{\text{eff}}}=\lambda _{0}^{2}/d$ , where d is the film thickness and ${{\lambda}_{0}}=\sqrt{{{m}^{\ast}}{{c}^{2}}/\left(4\pi n_{s}^{\ast}{{\left({{e}^{\ast}}\right)}^{2}}\right)}=\mathcal{O}\left({{10}^{4}}~\overset{\circ}{\mathop{\text{A}}}\,\right)$ is the London penetration depth. The Ginzburg–Landau coherence length is

Equation (190)

which is the length scale over which $|\Psi\left(\mathbf{r}\right)|$ varies so that $\Psi\left(\mathbf{r},z\right)$ is constant over the film thickness. Since one is interested in the temperature range over which phase fluctuations are important, it is clear that the parameter $r(T)\ll 0$ so that $|\Psi\left(\mathbf{r}\right)|\,=\mathcal{O}(1)$ since the phase is not defined otherwise. Thus, one can assume that the temperature $T\ll {{T}_{0}}$ , the Ginzburg–Landau critical temperature. The system is effectively two dimensional when the film thickness $d\ll {{\xi}_{0}}$ since we can integrate (189) over z to obtain an effective 2D theory with $|{{\Psi}_{0}}{{|}^{2}}=n_{s}^{0}$ , the Cooper pair areal number density, which is taken to be constant except at vortex cores. The vortex core has radius $\mathcal{O}\left({{\xi}_{0}}\right)$ which is the smallest length scale of the system.

When $\mathbf{H}=0$ , these considerations lead to the free energy $\mathcal{F}\left(\theta,\mathbf{A}\right)$

Equation (191)

From the extremal equation $\delta \mathcal{F}/\delta \mathbf{A}=0$ , we obtain

Equation (192)

where e*  =  2e and ${{m}^{\ast}}=2{{m}_{e}}$ are the charge and mass of a Cooper pair. To see the effect of the field $\mathbf{B}\left(\mathbf{r},z\right)$ due to a set of vortices in the superconducting plane at z  =  0, one follows the analysis of Pearl [237] and Nelson [238] by assuming the superconducting current is only in the z  =  0 plane so that

Equation (193)

where the flux quantum ${{\phi}_{0}}=\pi \hslash c/e$ , ${{\mathbf{A}}_{2d}}\left(\mathbf{r}\right)=\mathbf{A}\left(\mathbf{r},0\right)$ , and the transverse penetration depth ${{\lambda}_{\bot}}={{m}_{e}}{{c}^{2}}/\left(8\pi {{n}_{s}}{{e}^{2}}\right)$ . Now one can use the standard equations of magnetostatics to solve for the current due to a vortex of unit strength at $\mathbf{r}=0$ so that

Equation (194)

since $\nabla \theta \left(\mathbf{r}\right)=2\pi \left(\mathbf{\hat{z}}\times \mathbf{r}\right)/{{r}^{2}}$ due to the vortex at $\mathbf{r}=0$ and ${{\mathbf{A}}_{2d}}\left(\mathbf{r}\right)\equiv \mathbf{A}\left(\mathbf{r},z=0\right)$ . This can be solved by taking Fourier transforms, with the result [237, 238]

Equation (195)

which translates into the vortex–vortex interaction behaving as $\text{ln}r$ for $r<{{\lambda}_{\bot}}$ and as 1/r2 for $r\gg {{\lambda}_{\bot}}$ .

It is very important to note that the onset of superconductivity in 2D is due to vortex type phase fluctuations which requires a pre-existing condensate or pairing of electrons. This phenomenon is described by a Ginzburg–Landau free energy functional of (189) in which the condensate forms at the mean field temperature T0 when r(T0)  =  0. Even at T  <  T0, although there is a finite condensate or BCS pairing, $|{{\Psi}_{0}}{{|}^{2}}>0$ , there is not necessarily any superconductivity. This requires phase coherence or a finite renormalized superfluid density, just as in the neutral superfluid 4He which is controlled by vortices. The onset of phase coherence or true superconductivity in the thermodynamic limit of infinite system size $L\to \infty $ occurs when the experimentally measurable kinetic inductance ${{L}_{K\Box }}(T)={{m}_{e}}/\left(2{{n}_{s}}{{e}^{2}}\right)$ which is related to the transverse penetration depth

Equation (196)

In the absence of an external magnetic field $\mathbf{H}=0$ , the condition for the onset of superconductivity is given by the analogue of (46), [240, 241]

Equation (197)

where ${{T}_{\text{c}}}$ is the temperature at which phase coherence first appears

From this discussion, we see that there can be no true phase transition in a superconducting film because of the finite penetration depth ${{\lambda}_{\bot}}(T)$ for $T\leqslant {{T}_{\text{c}}}$ [19]. However, since ${{\lambda}_{\bot}}(T)$ can be $\mathcal{O}\left({{10}^{-2}}\right)$ m which is larger than most experimental systems so that we can regard most superconducting films at finite temperature when quantum effects are unimportant as described by the class of Hamiltonians

Equation (198)

for a uniform 2D superconducting film, where $n_{s}^{0}(T)$ is the bare superfluid number density and 2m the Cooper pair mass. In a clean uniform film, the mean field Cooper pair formation temperature T0 and the phase coherence onset temperature ${{T}_{\text{c}}}<{{T}_{0}}$ are rather close and effects due to these tend to become mixed together. It is known that Cooper pairing alone does not imply superconductivity which requires global phase coherence. Thus, to study superconductivity in a thin film, it is important that T0 and ${{T}_{\text{c}}}$ are well separated. This can be accomplished by a granular film with a large resistance RN in the normal state [242247] or by coupling the grains by weak links [248] as in arrays of Josephson junctions [249].

When the external magnetic field Hz  =  0, the superconducting film should be like a finite size 2D planar rotor model with $L={{\lambda}_{\bot}}$ . It was pointed out [240, 241] that the vortex unbinding mechanism [19] could be very important in thin superconducting films provided the transverse penetration depth ${{\lambda}_{\bot}}$ was very large which could be realized in very thin films with a large normal sheet resistance ${{\lambda}_{\bot}}\propto {{R}_{\Box }}$ . When ${{R}_{\Box }}>{{10}^{3}}\Omega $ , there is a reasonable separation between the pair formation and KT temperatures ${{T}_{0}}>{{T}_{\text{KT}}}$ which is essential for the experimental observation of vortex unbinding. They proposed [240] that ${{T}_{\text{KT}}}\approx 2.18{{T}_{0}}{{R}_{\text{c}}}/{{R}_{\Box }}$ but, on a more fundamental level [249, 250],

Equation (199)

where the vortex fugacity is ignored, which fits the data much better.

In a uniform superconducting film in a finite magnetic field B normal to the plane of the film there will be a hexagonal lattice of vortices each containing one flux quantum ${{\phi}_{0}}=\hslash c/2e$ which form a stable triangular lattice of lattice spacing a0 where [251, 252]

Equation (200)

where n is the areal vortex density due to the field. In the presence of arbitrarily weak pinning of vortices, the flux lattice will be pinned for $T<{{T}_{\text{m}}}$ , the lattice melting temperature, and vortices will be free to move when $T>{{T}_{\text{m}}}$ so that the onset of superconductivity can be identified with the vortex lattice melting temperature ${{T}_{\text{m}}}$ [252].

Equation (201)

and $\mu (T)=0$ when $T>{{T}_{\text{m}}}$ . In (201), the elastic constants should be interpreted as the renormalized ${{\mu}^{\text{R}}}(T)$ and ${{\lambda}^{\text{R}}}(T)$ . The phase diagram of a 2D superconductor in the (H, T) plane is sketched in figure 31 assuming that the transition is caused by vortex lattice melting [252].

Figure 30.

Figure 30. Phase diagrams for 2D systems of elongated molecules in the L−1, T plane where L is the shorter of the experimental length scale or ${{\xi}_{+}}(T)$ . The dashed lines are cross-overs between different behaviors at $L={{\xi}_{i}}(T)$ (see text). The left figure is for type I and the right is type II melting. Reprinted from [202] with permission. Copyright 1981 Elsevier.

Standard image High-resolution image
Figure 31.

Figure 31. Theoretical phase diagram in the (H, T) plane for a uniform thin film superconductor in a magnetic field H with vortex solid, hexatic and fluid phases. The flux lattice melting curve (TM) is the solid line and the hexatic-fluid transition (TH) is the dashed line. The H  =  0 BKT transition is at ${{T}_{\text{c}}}$ . The bulk upper critical field Hc2(T) is sketched as the dotted line terminating at the bulk BCS temperature Tc0. The lower critical field Hc1(T) is so small that, on the scale of the figure, it is indistinguishable from H  =  0. Reprinted from [252] with permission. Copyright 1980 Elsevier.

Standard image High-resolution image

In the most important field regime, $\xi _{\text{GL}}^{-2}\gg B/{{\phi}_{0}}\gg \lambda _{\bot}^{-2}$ , the bare ${{\lambda}_{0}}=\infty $ because of the long range vortex–vortex interactions and the bare shear modulus [253]

Equation (202)

where $0.4\leqslant C\leqslant 0.75$ [252]. When $T>{{T}_{\text{m}}}$ , theory says that there should be a hexatic fluid up to Ti but this is not relevant as there is no known probe coupling to a vortex hexatic. Of course, in any real system, there will be pinning of vortices by some random pinning sites [252] which will not be discussed in this review. However, it should be noted that the assumed KTHNY melting of the vortex lattice is controversial. Even the existence of a 2D flux lattice has been questioned [254]. Later work [255261] showed that, although numerical estimates of ${{\mu}_{\text{R}}}(T)$ are consistent with the KTHNY scenario, vortex lattice melting proceeds by a weakly first order transition so that ${{\mu}_{\text{R}}}(T)$ is very close to the KTHNY value.

There has been much interest in trying to apply the theory of 2D vortex lattice melting to thin films of YBCCO in a transverse magnetic field. This has met with only partial success [262264] as vortex lattice melting does not appear to be the mechanism for the destruction of superconductivity [265, 266]. Vortex pinning by random impurities dominates [267] so that the system is better described as a moving glass phase of a current driven moving elastic lattice [268271]. However, the basic ideas of dislocations and disclinations in a 2D vortex lattice have been verified by direct observation [272].

3.7. Superconducting arrays

A regular array of superconducting grains coupled by almost identical SNS or SIS junctions can be made with up to about 106 grains with modern lithographic deposition techniques [273] and almost any desired geometry and coupling constant arrangement can be constructed. Also, when an external magnetic field is applied, a frustrated XY model can be obtained, described by Coulomb gas Hamiltonian [274]

Equation (203)

where the first term is the charging energy with Cij the capacitance matrix, ${{q}_{i}}=2e{{n}_{i}}$ with ni the number of Cooper pairs on grain i, and ${{Q}_{i}}=2e{{q}_{\text{ex}}}=\sum_{j}{{C}_{ij}}V_{j}^{\text{ex}}$ is the charge on grain i induced by the external gate voltage. The capacitance matrix Cij has diagonal elements ${{C}_{ii}}={{C}_{0}}+zC$ where z is the number of nearest neighbors and Cij  =  −C for i, j nearest neighbors as shown in figure 32. ${{E}_{\text{J}}}$ is the Josephson coupling energy between adjacent grains and ${{A}_{ij}}=(2e/\hslash c)\int_{i}^{j}\mathbf{A}\centerdot \text{d}\mathbf{l}$ . The magnetic frustration is the flux f penetrating a plaquette given by $f=(1/2\pi )\sum_{\text{P}}{{A}_{ij}}$ where $\sum_{\text{P}}$ means the sum over the bonds round the plaquette in a counter clockwise direction. Quantum mechanical effects result from the non commuting operators $\left[{{\theta}_{i}},{{n}_{j}}\right]=\text{i}{{\delta}_{ij}}$ which become important at $T=\mathcal{O}{{\left({{10}^{-3}}\right)}^{{}^\circ}}K$ when the capacitances C0, C are small so the grain charging energy ${{E}_{\text{c}}}$ is comparable to the Josephson energy ${{E}_{\text{J}}}$ .

Figure 32.

Figure 32. Left: Reprinted from [309] with permission. Copyright 1991 Elsevier. Showing the state of understanding of the FFXY model in 1991. Right: Phase diagram of the RSOS Ising model with A  =  B dual to the FFXY model. Note the sliver of intervening Ising ordered phase for all values of C. Reprinted from [313] with permission. Copyright 1997 Elsevier.

Standard image High-resolution image

Many experiments are done on arrays of large capacity superconducting grains on a lattice coupled either by the proximity effect [248] or arrays of Josephson junctions with ${{E}_{\text{C}}}\ll {{E}_{\text{J}}}$ [249]. These are best described by a lattice model of the form

Equation (204)

where ${{A}_{ij}}=\int_{{{\mathbf{r}}_{i}}}^{{{\mathbf{r}}_{j}}}\text{d}\mathbf{r}\centerdot \mathbf{A}\left(\mathbf{r}\right)$ and charging effects can be ignored.

A junction array in a uniform magnetic field is a realization of an interesting class of models which, in the Coulomb gas representation is

Equation (205)

where i labels the center of the ith plaquette on the dual lattice, ${{n}_{i}}=0,\pm 1,\cdots $ and $0\leqslant {{f}_{i}}\leqslant 1/2$ the magnetic flux ${{f}_{i}}=2\pi /{{\phi}_{0}}\mathop{\oint}^{}\text{d}\mathbf{l}\centerdot \mathbf{A}$ penetrating the ith plaquette. The interaction Gij is the screened Coulomb interaction, ${{G}_{ij}}\sim \text{ln}\left(|{{\mathbf{R}}_{i}}-{{\mathbf{R}}_{j}}|/a\right)$ for $|{{\mathbf{R}}_{i}}-{{\mathbf{R}}_{j}}|<{{\lambda}_{\bot}}$ and ${{G}_{ij}}\sim \left({{a}^{2}}/|{{\mathbf{R}}_{i}}-{{\mathbf{R}}_{j}}{{|}^{2}}\right)$ when $|{{\mathbf{R}}_{i}}-{{\mathbf{R}}_{j}}|>{{\lambda}_{\bot}}$ . Even in this classical limit, these systems provide realizations of a great variety of models in 2D and have been the subject of intense activity in the 1980's. The models become even more interesting, and difficult, at very low T when quantum effects play a vital role [274282].

The systems with a Hamiltonian of (205) represent a large class of statistical mechanical models. When the frustration fi  =  0 in zero applied magnetic field, the system on a square lattice becomes a 2D planar rotor model when the screening length ${{\lambda}_{\bot}}\to \infty $ . In a junction array one can make this larger than the array size, so the junction array is a realization of the planar rotor model. In the phase representation the Hamiltonian of an array is [283],

Equation (206)

Because of the arbitrariness in the vector potential, for the fully frustrated case of f  =  1/2, one can choose Aij  =  0 on horizontal bonds and ${{A}_{ij}}=\pi $ on alternating vertical bonds. The system is mapped into one with ferromagnetic bonds of strength J on horizontal bonds and alternating vertical strength J ferromagnetic and strength $\eta J$ antiferromagnetic bonds [284]. The phase diagram in the $\eta,T$ plane from MC simulations shows that, away from the isotropic point $\eta =1$ , there is an Ising transition followed by an XY transition at higher T. The situation near the isotropic point $\eta =1$ is unclear but they seemed to occur simultaneously [284].

About the same time as this investigation was being done, there was much interest in the isotropic fully frustrated array with f  =  1/2 [283303]. It was well understood that the isotropic fully frustrated system has both chiral order and phase coherence at sufficiently low temperature and that phase coherence required chiral, Ising, long range order, but not vice-versa. Thus the main question was whether the XY transition happened at a lower temperature than the Ising transition or whether the two transitions happened simultaneously and, if so, what is the nature of this transition? The problem is analytically intractable and there have been many simulations over the years, mostly inconclusive with a few exceptions.

The problem was discussed in [283] where it was explained that the renormalized helicity modulus in the XY ordered phase must obey $\Upsilon (T)/{{k}_{\text{B}}}T\geqslant 2/\pi $ and will drop discontinuously to zero at $T={{T}_{\text{KT}}}$ with a jump to zero $\Delta \Upsilon /{{k}_{\text{B}}}{{T}_{\text{KT}}}\geqslant 2/\pi $ . There are two possibilities: (i) $\Delta \Upsilon /{{k}_{\text{B}}}{{T}_{\text{KT}}}=2/\pi $ at some ${{T}_{\text{KT}}}<{{T}_{I}}$ or (ii) as $T\to T_{I}^{-}$ , there is a non-universal jump $\Delta \Upsilon /{{k}_{\text{B}}}T>2/\pi $ . At first sight, these alternatives could be readily distinguished by simulation, but this hope was rapidly dashed because either the system sizes were too small to resolve the two transitions or they could be resolved with ${{T}_{\text{KT}}}<{{T}_{I}}$ but the jump in $\Upsilon (T)$ at ${{T}_{\text{KT}}}$ we larger than the universal value of $2/\pi $ . This situation led to the suggestion that the XY-Ising model [287] would be easier to simulate where the Hamiltonian is

Equation (207)

where ${{s}_{i}}=\pm 1$ represents the chirality of a plaquette. However, in the end it turns out that despite the short range interactions of the XY-Ising model, in the interesting region of $C\leqslant 0$ and A  =  B, simulations of the model of (207) turn out to be just as difficult as the original representation.

The derivation of (207) for frustrated systems was done in a series of papers [287, 288, 304310] by a set of approximate steps based on symmetry and renormalization group arguments and missed some of the short distance physics. For example, for an isotropic fully frustrated junction array [309], A  =  B so that there is no phase stiffness across an Ising domain wall with ${{s}_{i}}{{s}_{j}}=-1$ when i, j represent sites on opposite sides of domains of opposite chirality. In the original system, described by (206) with fi  =  f  =  1/2, there is a finite phase stiffness across domain walls. However, in an important paper, Korshunov [311] showed that domain wall kinks unbind at ${{T}_{\text{K}}}<{{T}_{\text{V}}}$ , the vortex unbinding or phase coherence temperature. It was also shown that ${{T}_{\text{V}}}<{{T}_{\text{DW}}}$ , the Ising critical temperature. These results were verified numerically [312]. This establishes, in an isotropic fully frustrated system, (i) phase coherence is destroyed at ${{T}_{\text{V}}}<{{T}_{I}}$ and (ii) the domain walls roughen at ${{T}_{\text{K}}}<{{T}_{\text{V}}}$ so that the XY-Ising model of (207) describes the long distance behavior near continuous transitions of (206). It is a reasonable model of a fully frustrated isotropic junction array on a square or triangular lattice. This is contrary to the findings in [289] where a difference was found between a square and triangular lattice when f  =  1/2.

There is yet another representation obtained by taking the dual of the XY part which yields the RSOS-Ising model on a square lattice with

Equation (208)

where the nearest neighbor sites  <i, j >  are on the original square lattice and $<\text{I},\text{J}>$ are the corresponding sites on the dual lattice. ${{h}_{I}}=0,\pm 1$ , ${{s}_{i}}=\pm 1$ and $\tilde{A}$ and $\tilde{C}$ are given in terms of A and C of (207), [313]. There is a constraint on the height variables hI such that a step in h is forbidden to cross a domain wall in s [313]. Another duality transformation on the si gives exactly the coarse grained model for CsCl(001) facets [314]. The XY-Ising model of (207) describes the critical behavior of a surprising number of models, including the 4-state Potts model and also contains the much discussed points where Ising and XY degrees of freedom are simultaneously critical at $A-B=\pm {{x}_{\text{c}}}$ where the Ising and XY transition lines cross, as shown in figure 32, center. There has been some discussion that new critical behavior might be found in isotropic systems at a transition in a new universality class between a fully ordered to a disordered phase [296, 309, 315317]. This was later shown not to happen by Olsson by large scale simulations which were sufficiently accurate to distinguish the two transitions [318, 319]. The phase diagrams of figure 32 summarise my understanding of these fully frustrated planbar rotor models.

Finally, it seems that some issues still remain in the isotropic fully frustrated system despite all the effort expended. In a relatively recent study of the FFXY model and the coupled Ising XY model (207) on the A  =  B line by large scale simulations with up to 106 sites on a square lattice [320], there is agreement with [313] for a wide range of the parameter A and C but disagreement when C  <  −5 when the Ising and XY transition lines seem to merge into a single line at a multicritical point at a critical value C* with  −5  >  C*  >  −7 [320]. In the RSOS Ising representation, the two critical lines do not merge into a single line, but just become very close together [313]. This disagreement is now purely philosophical and might be resolved by studying the full XYI model of (207) in the space of all parameters A, B, C. It is unlikely, but not impossible, that a single multicritical point will emerge for $C\ll 0$ with $A-B\to 0$ but such a scenario will require massive computations. The issue of systems with nearby Ising and XY transitions also arises in the competition between surface reconstruction and roughening in (1 1 0) facets of fcc crystals which are described by a two component body centered solid on solid model or a staggered 6-vertex model [321324]. Unfortunately, this is also insoluble but is amenable to accurate transfer matrix studies.

There are few experiments on these fully frustrated systems because the XY and Ising transitions are too close to resolve in an isotropic junction array. However, they become quite well separated in an anisotropic array [284, 305] but there are some technical issues in fabricating an array with the required anisotropy in the coupling strengths. The helicity modulus in an isotropic array with Hamiltonian

where $\mu =x,y$ for a square lattice. The measurable helicity moduli are [305] near the low T Ising transition reflects the Ising singularity [305]

Equation (209)

This very weak singularity was observed in a custom built junction array [325] by measuring the complex sheet impedance $Z(\omega T)=R(\omega,T)+\text{i}\omega L(\omega,T)$ and observing a small peak in the resistivity at ${{T}_{I}}<{{T}_{\text{KT}}}$ well below the BKT temperature obtained from the universal relation $L_{\Box }^{-1}\left({{T}_{\text{KT}}}\right)=\left(8\pi /{{\phi}_{0}}\right){{k}_{\text{B}}}{{T}_{\text{KT}}}$ . The custom built junction array and the peak in the resistivity are shown in figure 33.

Figure 33.

Figure 33. Left: a small piece of the anisotropic array with alternating weak (WB) and strong coupling (SB) rows. Right: The peak in R at ${{T}_{I}}<{{T}_{\text{KT}}}$ . Reprinted from [325] with permission. Copyright 2002 Elsevier.

Standard image High-resolution image

Quantum effects become important at very low temperatures and when the charging and Josephson energies are comparable ${{E}_{\text{c}}}/{{E}_{\text{J}}}=\mathcal{O}(1)$ . This requires very small junctions as discussed in a comprehensive recent review [274]. The partition function corresponding to the Hamiltonian of (203) is [275]

Equation (210)

where $\beta =1/{{k}_{\text{B}}}T$ and $\overset{\centerdot}{\mathop{\phi}}\,=\text{d}\phi /\text{d}\tau $ . The first and second terms are charging energies in terms of the voltages at the lattice sites. To include dissipative tunneling, the action includes a term

Equation (211)

where, for Ohmic dissipation

Equation (212)

The dissipation mechanism might be normal electron tunneling by discrete charge transfer and

Equation (213)

which is $4\pi $ periodic. Other mechanisms which yield quadratic forms of $F[\phi ]$ [274] and may force the phase $\phi (\tau )$ to be continuous. If the dissipation is by discrete charge transfer, the path integral implies a sum over winding numbers so that

Equation (214)

because the charges on the grains are integer multiples of 2e [274].

To obtain the dual Coulomb gas action for a Josephson junction array, one expresses the array partition function in terms of a sum over charge and vortex configurations as $Z=\sum_{q,v}\text{exp}\left[-S(q,v)\right]$ by [274]

Equation (215)

where the charge ${{q}_{i}}(\tau )$ and the vorticity ${{v}_{i}}(\tau )$ are integer valued functions of imaginary time τ. Gij is the interaction between two vortices separated by rij/a lattice spacings with Ea the short distance contribution. Provided the capacitance to ground ${{C}_{0}}\ll C$ , the capacitance between neighboring grains, the interaction between charges has the same form as between vortices

Equation (216)

so that, from (215), the charges and vortices are dual with a self dual point at ${{E}_{\text{J}}}/{{E}_{\text{C}}}=2/{{\pi}^{2}}$ which is exact when C0  =  0 and in the absence of the spin wave term $\overset{\centerdot}{\mathop{q}}\,G\overset{\centerdot}{\mathop{q}}\,$ in (215). This term is irrelevant for the statics but not for dynamics. These considerations lead to a phase diagram as shown in figure 34, right, which should be compared with the experimental phase diagram of figure 34, center.

Figure 34.

Figure 34. Reprinted from [274] with permission. Copyright 2001 Elsevier. The figure to the left is a sketch of a piece of an array with the parameter C0 and C shown, the middle figure is the measured phase diagram for f  =  0 for a square (solid squares) and a triangular (solid triangles) showing the S-I transition for ${{E}_{\text{c}}}/{{E}_{\text{J}}}\sim 1.7$ . The solid line is a guide to the eye and the dotted line on the superconducting side is from a calculation [280, 281].

Standard image High-resolution image

There are some interesting predictions arising from these considerations related to the experimental observation that, at the border of the superconducting -insulator transition, the 2D system is metallic with a resistivity ${{\rho}^{\ast}}(T=0)\approx {{\rho}_{Q}}=h/{{(2e)}^{2}}$ , the quantum of resistance. An oversimplified argument is in [326] based on vortex flow and duality. The Josephson relation for the voltage V across a junction and the current I are

Equation (217)

At the self dual point of (215), we have $\langle q\rangle =\langle v\rangle $ , which means that for every Cooper pair crossing a square, one vortex crosses perpendicular to the Cooper pair. Thus, at the T  =  0 SI transition the system is metallic with resistivity ${{\rho}^{\ast}}=h/\left(4{{e}^{2}}\right)$ . Of course, the self duality is not exact but one might expect that ${{\rho}^{\ast}}=\mathcal{O}\left({{\rho}_{Q}}\right)$ , especially as the self dual point is a universal fixed point in the RG sense. Some explicit calculations of ${{\rho}^{\ast}}$ have been performed, mostly as a 1/N expansion [326328]. In the exactly soluble $N=\infty $ limit, the universal resistivity of a superconducting array ${{\rho}^{\ast}}(0)=8{{R}_{Q}}/\pi $ in zero applied field [327] and, in the fully frustrated case of f  =  1/2, ${{\rho}^{\ast}}(1/2)={{\rho}^{\ast}}(0)/2=4{{R}_{Q}}/\pi $ [328]. Measurements of the T  =  0 universal resistivity [329] are not easy.

4. Dynamics

As mentioned earlier, to interpret experiments like the torsional oscillator measurements of the superfluid density in a 2D 4He film, one must extend the static theory discussed to finite frequency ω. This is because the quantity called the superfluid density is the $q=\omega =0$ limit of the linear response function ${{\rho}_{s}}(q,\omega )$ where ${{g}_{s}}(q,\omega )={{\rho}_{s}}(q,\omega ){{v}_{s}}(q,\omega )$ is the superfluid momentum density and ${{v}_{s}}(q,\omega )$ the superfluid velocity. The response function ${{\rho}_{s}}(q,\omega )$ happens to have the dimensions of mass per unit area but has to be measured dynamically by, for example, Bishop and Reppy's torsional oscillator experiment [67, 68] which must be done at $\omega >0$ or a third sound measurement at q  >  0 and $\omega >0$ . The quantity ${{\rho}_{s}}$ for which exact calculations can be done is ${{\rho}_{s}}\left(q,\omega =0\right)$ which is inaccessible to experiment. To relate theory [19, 20, 56, 63] to experiment [36, 37, 6769], the dynamical extension of the static theory is needed.

This was done at the same time as the key Bishop–Reppy experiment by a group at Cornell [76, 330, 331]. Following the early work of Hall and Vinen [332336] by balancing Magnus and drag forces, the position ${{\mathbf{r}}^{\nu}}$ of the νth vortex of strength ${{n}_{\nu}}=\pm 1$ obeys [76]

Equation (218)

ignoring boundaries.

By writing the Fokker–Planck equation for the N vortex distribution function ${{\Gamma}_{N}}\left({{\mathbf{r}}^{\nu}},t\right)$ corresponding to (218) [76, 330, 331] the scaling equations for the various parameters can be found. The scaling equations to $\mathcal{O}\left(\,{{y}^{2}}\right)$ for K(l) and the fugacity y(l) are the same as for the statics and, for the diffusion constant D(l) and the parameter C(l) are

Equation (219)

The renormalized vortex diffusion constant ${{D}_{\text{R}}}(T)=D\left(l=\infty \right)$ $={{D}_{\text{R}}}\left({{T}_{\text{c}}}\right)\left(1+b{{t}^{1/2}}\right)$ with ${{D}_{\text{R}}}\left({{T}_{\text{c}}}\right)>0$ so that ${{D}_{\text{R}}}(T)$ has a cusp at ${{T}_{\text{c}}}$ . As pointed out in [76] this has little effect so that the dynamics is described by diffusive motion of the vortices supplemented by using renormalized values of K(T). Experimentally, the vortex diffusivity D does have some dependence on T [337, 338] and seems to increase as $T\to T_{\text{c}}^{-}$ , which is the opposite of the theoretical prediction [330, 331]. To my knowledge, this discrepancy has not been resolved and it is not known if it is significant. When the frequency ω and the average superfluid velocity ${{\mathbf{u}}_{s}}$ are small, one immediately sees that the average superfluid velocity in a film of width W and length L decays by the motion of vortices perpendicular to ${{\mathbf{u}}_{s}}$ as

Equation (220)

A uniform flow of superfluid relative to the substrate decays slowly to zero because superfluidity in 2D is only metastable and vortex pairs are driven apart over the free energy barrier to become free as if the system is above ${{T}_{\text{c}}}$ .

Ambegaokar et al [76] also noted that the motion of vortices on an oscillating substrate is equivalent to the diffusive motion of 2D charges $\pm {{q}_{0}}$ where $q_{0}^{2}=\pi \rho _{s}^{0}\hslash \,/\,m$ between capacitor plates in an external electric field with ${{q}_{0}}{{\mathbf{E}}_{\text{ext}}}=2\pi \rho _{s}^{0}(\hslash /m)\mathbf{\hat{z}}\times {{\mathbf{v}}_{n}}$ . In linear response theory, the the time dependent behavior of the system, for $|{{\mathbf{v}}_{n}}|\ll 1$ and $\omega \ll 1$ is encoded in a dielectric function $\epsilon (\omega )$ as a function of frequency ω defined by

Equation (221)

Using this capacitor analogy, it is not difficult to see that, when the substrate oscillates as ${{v}_{n}}(t)={{v}_{n}}{{\text{e}}^{\text{i}\omega t}}$ , the power dissipated per unit area is

Equation (222)

and, provided the oscillating system has no dissipation other than in the 4He film of area A, the Q and the shift $\Delta P$ of the oscillator period P are

Equation (223)

Thus, most important dynamical behavior is encoded in the the dielectric function $\epsilon (\omega )$ defined in (221) which can be estimated in different regimes.

In the Coulomb gas language, a vortex of unit circulation can be regarded as a particle with charge ${{q}_{0}}={{\left(2\pi \rho _{s}^{0}\right)}^{1/2}}\hslash /m$ in an electric field $\mathbf{E}$ with ${{q}_{0}}{{\mathbf{E}}_{\text{ext}}}=2\pi \rho _{s}^{0}(\hslash /m)\mathbf{\hat{z}}\times {{\mathbf{v}}_{n}}$ . The energy U(r) of a pair of vortices of circulation $n=\pm 1$ and separation r is [19, 63, 76]

Equation (224)

where $\tilde{\epsilon}(r)={{K}_{0}}{{K}^{-1}}(r)$ , defined in (33), accounts for the vortex pairs of separation less than r. The contribution of all bound vortex pairs to the dielectric function $\epsilon (\omega )$ for $T<{{T}_{\text{c}}}$ when ${{\xi}_{-}}(T)\ll {{r}_{\text{D}}}=\sqrt{(14D/\omega )}$ , where D is the vortex diffusion constant, is [68, 76, 331]

Equation (225)

Nothing very dramatic happens as $T\to T_{\text{c}}^{-}$ but, when $T>{{T}_{\text{c}}}$ , vortex pairs with separation $r>{{\xi}_{+}}(T)$ are free and the individual vortices diffuse freely in the field $\mathbf{E}$ and (225), using the Coulomb gas analogy,

Equation (226)

Here, the analogy of the motion of charges between capacitor plate is used [76] and the estimate of the free vortex density ${{n}_{f}}=F\xi _{+}^{-2}$ where F is a constant $\mathcal{O}(1)$ [76].

Putting everything together, one arrives at expressions for $\epsilon (\omega )$ [68, 339] in terms of experimentally measurable quantities and a direct comparison of theory and experiment [68, 339, 340] can be made as shown in figure 35. However, other samples do not fit the predictions of theory as well as this which is believed to be due to substrate inhomogeneity [339]. The agreement is remarkable and the data also gives a value for the static exponent $\eta \left(T_{\text{c}}^{-}\right)=0.25\pm 15\%$ which again agrees well with the theoretical value of 0.25. This gives some confidence that the vortex theory [19, 20, 56] and the dynamical extension [76] is a good description of the physics of thin 4He and superconducting films, in the linear response regime.

AHNS [76] discuss the effects of vortices on the third sound modes in thin superfluid films, which are due to variations in the film thickness or mass density, [341346], using the hydrodynamic equations of Bergman [347, 348] supplemented by equations for the vortex density $N\left(\mathbf{r},t\right)$ and vortex current ${{\mathbf{J}}_{v}}\left(\mathbf{r},t\right)$

Equation (227)

where $\bar{S}$ is the film entropy per unit mass and f the van der Waals constant. To simplify this rather complicated problem but keeping the essential physics, one can ignore variations in temperature in the film and mass transport between film and vapor so that

Equation (228)

Now one can write down a complete set of equations for the film which are just like the Maxwell equations of electromagnetism [76]. Defining

Equation (229)

The last step is to absorb the effects of the bound vortex (charge) pairs into the dielectric constant ${{\epsilon}_{b}}\approx {{\epsilon}_{\text{c}}}=\rho _{s}^{0}/{{\rho}_{s}}\left(T_{\text{c}}^{-}\right)$ , so that the free vortex density ${{N}_{\text{free}}}\left(\mathbf{r},t\right)={{N}_{+}}\left(\mathbf{r},t\right)-{{N}_{-}}\left(\mathbf{r},t\right)\equiv \delta N\left(\mathbf{r},t\right)$ and

Equation (230)

The last step is to use the Langevin equation for ${{\overset{\centerdot}{\mathop{\mathbf{r}}}\,}_{i}}$ to express the free vortex current ${{\mathbf{J}}_{v,\text{free}}}$ in terms of ${{\mathbf{v}}_{s}}$ and one finds [76], in a linear approximation

Equation (231)

To find the frequencies of the modes, one takes Fourier transforms of everything in (230) and write ${{\mathbf{v}}_{s}}\left(\mathbf{q},\omega \right)={{\mathbf{v}}_{\text{L}}}+{{\mathbf{v}}_{\text{T}}}$ , where ${{\mathbf{v}}_{\text{L},\text{T}}}$ are the longitudinal and transverse components of ${{\mathbf{v}}_{s}}$ . ${{\mathbf{v}}_{\text{L}}}\left(\mathbf{q},\omega \right)$ couples to $m\left(\mathbf{q},\omega \right)$ and the eigenfrequencies of the longitudinal modes are solutions of [76]

Equation (232)

Thus, for $q\ll {{\gamma}_{0}}/2\sqrt{{{\epsilon}_{\text{c}}}g\rho _{s}^{0}}$ the eigenfrequencies are

Equation (233)

The transverse part of ${{\mathbf{v}}_{s}}$ relaxes as ${{\omega}_{0}}(q)=-\text{i}\left({{\gamma}_{0}}/{{\epsilon}_{\text{c}}}+D{{q}^{2}}\right)\approx $ $-\text{i}{{\gamma}_{0}}/{{\epsilon}_{\text{c}}}\sim -\text{i}\xi _{+}^{-2}$ [76].

These results can be used to define a measurable mass transport coefficient λ by [76]

Equation (234)

Comparing this with the relaxation rate ${{\omega}_{-}}$ from (233) one sees that $\lambda ={{D}_{h}}/{{g}^{2}}\sim \xi _{+}^{2}$ . Remarkably, measurements of λ [70, 71, 340] agree that $\lambda \sim \xi _{+}^{2}$ as $T\to T_{\text{c}}^{+}$ although the exponential dependence on $T-{{T}_{\text{c}}}$ differs somewhat from the torsional oscillator parameters [340]. When $T\leqslant {{T}_{\text{c}}}$ , third sound modes will propagate and have a dispersion $\omega (q)={{c}_{3}}q$ where c3 is the third sound velocity. From (232), for ${{q}^{2}}>\gamma _{0}^{2}/\left(4{{\epsilon}_{\text{c}}}g\rho _{s}^{0}\right)$ , there are two propagating but damped modes and, when $T<{{T}_{\text{c}}}$ , ${{\gamma}_{0}}=0$ , so that there are propagating modes for all q. AHNS [76] argue that, for $T<{{T}_{\text{c}}}$ , ${{\epsilon}_{\text{c}}}$ should be replaced by ${{\epsilon}_{b}}\left(\mathbf{q},\omega \right)=\epsilon (\omega )$ and ${{\mathbf{J}}_{v,\text{free}}}=0={{N}_{\text{free}}}$ so that

Equation (235)

As $q\to 0$ with $T-{{T}_{\text{c}}}<0$ and fixed

Equation (236)

where D3  >  0 is a constant and K(T) is the renormalized stiffness constant (53). When $T={{T}_{\text{c}}}$ , we have

Equation (237)

Thus, it follows that third sound propagates when $T<{{T}_{\text{c}}}$ for any wavenumber $q\ne 0$ and for $q>{{\gamma}_{0}}/\left(2\sqrt{{{\epsilon}_{\text{c}}}g\rho _{s}^{0}}\right)\sim \xi _{+}^{-2}$ when $T>{{T}_{\text{c}}}$ [76]. There is an excellent discussion of thermal conduction and third sound in 4He films by Teitel [349] where he attempts to fit the thermal resistance data using parameters from the simultaneous torsional oscillator dissipation data as shown in figure 36.

Figure 35.

Figure 35. The fractional period shift $\Delta P/P$ and dissipation Q−1 for a torsional oscillator due to a film of 4He adsorbed on a mylar substrate. The solid lines are from theory and the symbols are experimental. The dashed curve is from the static theory. Reprinted from [67] with permission. Copyright 1978 American Physical Society.

Standard image High-resolution image
Figure 36.

Figure 36. The period shift $2\Delta P/P$ , dissipation Q−1 [67, 68] and thermal resistivity. Reprinted from [349] with permission. Copyright 1982 Springer.

Standard image High-resolution image

Of course, it is very easy experimentally to get out of this regime into a regime where non linear effects dominate. In a torsional oscillator, large amplitude oscillations are all that is needed and a simpler situation is the decay of a uniform superfluid flow or of a persistent current in a superconductor. In this situation, the superfluid velocity ${{\mathbf{u}}_{s}}$ or current in a superconductor decays by the unwinding of the phase of the order parameter $\psi \left(\mathbf{r}\right)=|\psi \left(\mathbf{r}\right)|{{\text{e}}^{\text{i}\theta \left(\mathbf{r}\right)}}$ by the motion of vortices perpendicular to ${{\mathbf{u}}_{s}}$ .

The non linear effects have been studied by several workers [76, 339, 350, 351] who conclude that the decay of the mean superfluid velocity is due to the escape of vortices over the energy barrier at pair separation ${{r}_{\text{c}}}$ given by $E{{r}_{\text{c}}}\tilde{\epsilon}\left({{r}_{\text{c}}}\right)=2{{q}_{0}}$ . AHNS [76] show that the rate of production of free vortices by escape over the barrier is

Equation (238)

where R is the rate of escape over the barrier, the capture cross section ${{\sigma}_{s}}\sim {{r}_{\text{c}}}$ and ${{v}_{\bot}}$ the velocity of the vortices normal to ${{\mathbf{u}}_{s}}$ . The free vortex production rate $R=2D{{y}^{2}}\left({{l}^{\ast}}\right){{\text{e}}^{-4{{l}^{\ast}}}}$ [76] where l* depends on the two large length scales ${{r}_{\text{c}}}$ and ${{\xi}_{-}}(T)$ .

One obtains R in the two limits ${{r}_{\text{c}}}\gg {{\xi}_{-}}$ and ${{r}_{\text{c}}}\ll {{\xi}_{-}}$ [76]

Equation (239)

Since nf relaxes quickly it will be determined by the instantaneous value of the superfluid velocity so that

Equation (240)

When one applies a uniform super current density ${{\mathbf{J}}_{s}}$ there is dissipation caused by a flow of current induced vortices perpendicular to the direction of ${{\mathbf{J}}_{s}}$ from (218). From the Josephson relation [352, 353] a mean electric field E is generated by the vortex motion perpendicular to the super current ${{\mathbf{J}}_{s}}$

Equation (241)

where ${{v}_{\bot}}$ is the mean velocity of the vortices normal to ${{J}_{s}}\propto {{u}_{s}}$ , the superfluid velocity. Thus one obtains [351]

Equation (242)

where

Equation (243)

Thus, for infinitesmally small js, we have $V\sim {{I}^{\alpha (T)}}$ where $\alpha \left(T>{{T}_{\text{c}}}\right)=1$ corresponding to the expected Ohmic behavior, $\alpha \left({{T}_{\text{c}}}\right)=3$ and $\alpha \left(T<{{T}_{\text{c}}}\right)=1+\pi {{K}_{\text{R}}}(T)>3$ which agrees with experiment [247249] as shown in figure 37, left. However, in a more recent study on a thin film of high ${{T}_{\text{c}}}$ superconductor [354] of figure 37, right, the IV relation is also power law but this was used to argue for the absence of a KT transition in the high ${{T}_{\text{c}}}$ film because there is Ohmic behavior at $T\ll {{T}_{\text{c}}}$ . A commonly used criterion to estimate ${{T}_{\text{KT}}}$ from the IV relations of a thin film is to define ${{T}_{\text{KT}}}$ as the temperature at which $\alpha (T)=3$ . However, a quick glance at (242) shows that, as js increases, the IV characteristics all become power laws with $\alpha \approx 3$ . Also, when js is small enough, all the IV curves become Ohmic because of the finite screening length in charged superfluids [18, 19, 355]. For large currents ${{j}_{s}}/{{j}_{0}}\gg \xi _{-}^{-1}$ , from (242) we see that $E\sim {{I}^{3}}$ which seems to agree with the data of figure 37 for both t  <  0 and t  >  0 [355]. This follows because, when ${{j}_{0}}/{{j}_{s}}<{{\xi}_{-}}$ , (242) becomes

Equation (244)
Figure 37.

Figure 37. Left: IV characteristics of a ${{5.10}^{-4}}\times {{1.10}^{-4}}$ m, 100 Å thick film of indium/indium-oxide at H  =  0. This obeys the expected theory with a KT transition at 4.16 °K. Reprinted from [247] with permission. Copyright 1983 American Physical Society. Right: IV characteristics in a single unit cell film of $\text{YB}{{\text{a}}_{2}}\text{C}{{\text{u}}_{3}}{{\text{O}}_{7-\delta}}$ at $T\ll {{T}_{\text{c}}}$ showing Ohmic behavior at very small currents. The dashed line corresponds to $V\propto I$ . Reprinted from [354] with permission. Copyright 1996 Elsevier.

Standard image High-resolution image

Making a quick comparison of this with the IV characteristics in figure 37 it seems that they agree with (244) for the large values of ${{j}_{s}}/{{j}_{0}}$ in the upper left corner where all the IV characteristics appear to have almost the same slope of 3.

Early investigations of 2D superconductivity and the transition to the normal state were first discussed in [240, 242249]. The fits of experimental data to theory is, at fist sight, very convincing but some of the fits are spurious. The fitting of the experimental IV relation for $T\leqslant {{T}_{\text{c}}}$ is fine but the agreement with the resistivity $R\sim {{n}_{f}}\sim \text{exp}\left(-b/\sqrt{t}\right)$ [245, 247, 249, 356, 357] is outside the true critical region where the exponential form of the resistivity is not valid. Similar criticism can be made of a number of studies of the onset of superconductivity in films of $YB{{a}_{2}}C{{u}_{3}}{{O}_{7}}$ films [358366] which claimed KT behavior near onset, even for thick films.

More recent experiments on the dynamics of superconducting arrays of SNS coupled grains on a triangular lattice where vortex pinning by impurities is minimized show that the vortex dynamics has some surprising behavior [367] where it was found that the vortex mobility ${{\mu}_{v}}(\omega )$ vanished as the frequency $\omega \to 0$

Equation (245)

where ${{\mu}_{0}}={{\omega}_{0}}/2\pi q_{v}^{2}{{n}_{v}}$ with the vortex charge ${{q}_{v}}=\sqrt{\pi {{E}_{\text{J}}}}$ and ${{n}_{v}}=4f/\left(\sqrt{3}{{a}^{2}}\right)$ the areal vortex density.

This experimental observation agrees with the conjecture by Minnhagen [350] later justified [368370] who showed that, in the absence of any pinning by impurities, vortex dynamics become very sluggish as $\omega \to 0$ due to their coupling to spin waves. However, the results were somewhat puzzling until a study [371] demonstrated that both SIS and SNS junction arrays should show similar behavior and apparent differences can be explained by different sizes of the three regions. The dynamics of these superconducting arrays is determined by the competing length scales ${{\xi}_{+}}(T)$ and the probe length $r(\omega )=\sqrt{{{\Gamma}_{0}}/\omega}$ . The three regions are (i) $r(\omega )\gg {{\xi}_{+}}$ one recovers the expected Drude free vortex behavior in the hydrodynamic region with the vortex 'conductivity' ${{\sigma}_{1}}(\omega \to 0)\sim \xi _{+}^{2}$ . (ii) At intermediate scales $r(\omega )\sim {{\xi}_{+}}$ the ${{\sigma}_{1}}\sim \text{ln}\omega $ behavior [350, 367369] is recovered, while, at high frequencies, (iii) $r(\omega )\ll {{\xi}_{+}}$ , in a temperature regime from $T>{{T}_{\text{KT}}}$ down to T  =  0, one finds a scale dependent vortex damping with ${{\sigma}_{1}}(\omega )\sim {{\left(r(\omega )/\text{l}{{\text{n}}^{2}}\omega \right)}^{2}}\sim {{\left(\omega {{\left(\text{ln}\omega \right)}^{2}}\right)}^{-1}}$ corresponding to large vortex pairs moving in a viscous medium of smaller pairs. A finite free density of free vortices when the external field B  =  0 is obtained when $T>{{T}_{\text{KT}}}$ and, for $T\leqslant {{T}_{\text{KT}}}$ , when $B\ne 0$ when ${{n}_{v}}=2\pi f\sim \xi _{+}^{-2}$ where $f=\phi /{{\phi}_{0}}$ is the number of flux quanta per plaquette.

One non invasive probe of a 2D superconducting film or array is the two coil mutual inductance method [243, 247, 372379] which measures the voltage induced in the detection coil by the currents in the film induced by the current in the drive coil. This method measures the complex impedance per unit area at frequency ω, ${{Z}_{\Box }}(\omega )$ . Another method is even less invasive which is the flux noise spectrum $S(\omega )$ which is related to the impedance ${{Z}_{\Box }}$ by [370]

Equation (246)

where the geometrical factor F(q) is determined by the geometry of the detector coil. For a circular pick up coil of N turns of radius r, thickness b and height h above the superconducting film

Equation (247)

This can be related to correction to the mutual impedance $\delta {{Z}_{m}}(\omega )$ as measured by the two coil method

Equation (248)

Here, $\tilde{F}(q)$ is the geometrical factor for the drive coil, and, ignoring the difference between F(q) and $\tilde{F}(q)$ , one sees that $S(\omega )$ and $\delta {{Z}_{m}}(\omega )$ are linearly related

Equation (249)

Following the discussion in [370], when the height h of the coil is sufficiently large one can ignore the q dependence of ${{Z}_{\Box }}(q,\omega )$ and write ${{Z}_{\Box }}(q,\omega )=-\text{i}\omega {{L}_{\Box }}(\omega )+{{R}_{\Box }}(\omega )$ so that, for weak screening and ${{L}_{s}}(q)\ll {{L}_{\Box }}$ ,

Equation (250)

which is consistent with [379]. When $T\leqslant {{T}_{\text{KT}}}$ , ${{L}_{\Box }}(\omega )$ is almost constant, so that [370],

Equation (251)

which agrees with observation [380, 381]. In the strong screening limit, ${{L}_{\Box }}(\omega )\ll {{L}_{s}}(q)$ and it can be shown that the ω dependence depends very much on the geometry of the experiment [370]. For $r\gg h$ , one finds $S(\omega )\sim 1/\omega $ , the ubiquitous 1/f noise. However, it seems difficult to understand the observed 4 decade range of this $1/\omega $ noise [380, 381]. Noise measurements in films of YBCO [382384] also show that $S(\omega )\sim 1/\omega $ for ${{10}^{0}}<\omega <{{10}^{3}}$ which appears consistent with this discussion.

Perhaps this is not the full story but this work [370] is more successful in explaining experimental observations than most others. To obtain the $1/\omega $ behavior over 4 decades requires a ratio $r/h\sim {{10}^{4}}$ which is much larger than the experimental ratio. Thus, it seems that this aspect of the flux noise experiments is still awaiting explanation.

5. Cold atoms and conclusion

In the last decade the number of papers on cold atoms has exploded and there is renewed interest in topic such as Bose Einstein condensation (BEC) and superfluidity in two dimensional systems [385404]. From my simple minded viewpoint, in 2D the superfluid transition is the onset of quasi long range phase coherence for which a prerequisite is that the system is in its ground state or very close to the ground state. The BEC transition occurs at a lower T or density when one of the single particle states gets a macroscopic occupancy. It can therefore be described by the quantum mechanical analogue of the classical free energy functional [392]

Equation (252)

where g is the coupling parameter for the particle interactions, and $V\left(\mathbf{r}\right)$ is the harmonic laser trapping potential [390]. To trap the gas as a 2D pancake, the laser trapping potential has the form

Equation (253)

where ${{r}^{2}}={{x}^{2}}+{{y}^{2}}$ and the anisotropy parameter $\lambda ={{\omega}_{z}}/{{\omega}_{0}}\gg 1$ .

The interaction parameter g has nothing to do with the trapping potential $V\left(\mathbf{r}\right)$ which, at least for for small interaction strength g, is responsible for the condensate fraction ${{N}_{0}}/N=1-{{\left(T/{{T}_{0}}\right)}^{2}}$ where the BE condensation temperature is given by ${{k}_{\text{B}}}{{T}_{0}}=\hslash {{\omega}_{0}}\sqrt{N/\zeta (2)}$ in the 2D limit $\lambda =\infty $ . However, to my naive thinking, the quantum mechanical cold trapped atom literature seems to be very confused about the definition of the condensate density which should be different to the definition of the superfluid density. Some years ago, Leggett [386, 387] defined the global superfluid density fraction by considering N Bosons enclosed by a rotating circular surface of radius R with angular speed ω [386]. He defined the superfluid fraction

Equation (254)

where $\langle L\rangle $ is the expectation value of the mechanical angular momentum and ${{I}_{\text{c}}}=Nm/2{{R}^{2}}$ is the classical moment of inertia. Using rigorous but subtle arguments, he proved, among other things, that, in the thermodynamic limit, that fs(0)  =  1 for interacting Bosons and fs(0)  =  0 if the Bosons do not interact, thus establishing that BEC does not imply superfluidity.

It is well known that the Bose Einstein onset temperature T0 is finite for N non interacting Bosons in a 2D harmonic trap of radius R is [390]

Equation (255)

where n is the areal number density of particles. In an interesting paper [396], the relation of T0 and ${{T}_{\text{c}}}$ in a trapped gas was shown to be

Equation (256)

for typical experimental conditions where $\lambda =\sqrt{2\pi {{\hslash}^{2}}/m{{k}_{\text{B}}}T}$ is the 2D thermal wavelength so that ${{T}_{0}}>{{T}_{\text{c}}}$ and the density ${{n}_{\text{c}}}$ at ${{T}_{\text{c}}}$ is

Equation (257)

where $2\pi C\approx 380\pm 3$ [388, 389].

Somewhat surprisingly, a BKT transition is observed in a gas of 87Rb atoms trapped in a 2D optical trap [395]. It is expected that, when sufficiently cold, an isolated bosonic particle will condense into its ground state of zero momentum which is well known from elementary quantum statistical mechanics, and in general a system of interacting system of bosons will condense into its ground state with wavefunction $\psi \left(\mathbf{r}\right)$ . The experiment of [395] consists of optically trapping the bosons in two parallel planes at x1 and x2 separated by d with about 105 atoms per plane. These 2D Bose Einstein condensates (BEC) are allowed to equilibrate independently [393, 395]. The problem now is to detect the the algebraic correlations $\langle a_{\alpha}^{\dagger}\left(\mathbf{r}\right){{a}_{\alpha}}\left(\mathbf{0}\right)\rangle \sim {{\left({{\xi}_{h}}/r\right)}^{\eta (T)}}$ in each independent layer $\alpha =1,2$ .

The solution is to turn off the trapping potentials and to allow the two independent correlated clouds of particles to expand and merge into a single 3D cloud [395]. The two gases expand most rapidly perpendicular to the planes and interfere with each other. A probe laser beam parallel to the planes (in the y direction) to examine the interference pattern formed when the clouds overlap. Before the expansion, each plane of atoms is highly correlated and it is reasonable to assume that different planes are uncorrelated with each other since allowing the system to equilibrate ensures that the planes are independent of each other. At sufficiently low T, individual planes will have quasi long range phase order which is encoded in the 3D cloud after expansion in the density fluctuations which are detected by observing the interference fringes in the density–density correlation function as in figure 38 [394, 395]

Equation (258)

where ${{\mathbf{R}}_{j}}=\left({{\mathbf{r}}_{j}},{{z}_{j}}\right)$ , $\mathbf{r}=(x,y)$ . The operators ${{a}_{\text{tof}}}\left(\mathbf{r},z\right)=$ ${{a}_{1}}\left(\mathbf{r}\right){{\text{e}}^{\text{i}{{Q}_{1}}\left(z-{{Q}_{1}}t/2m\right)}}+{{a}_{2}}\left(\mathbf{r}\right){{\text{e}}^{\text{i}{{Q}_{2}}\left(z-{{Q}_{2}}t/2m\right)}}$ where ai is an operator of condensate i and ${{Q}_{1,2}}=m(z\pm d/2)/\hslash t$ where t is the time of flight during the free expansion of the atom clouds. After a bit of algebra, one finds that the correlation function in (258) contains an oscillatory piece with $Q=md/\hslash t$ [394, 395]

Equation (259)

when the two planes of condensate are independent before the expansion occurs [394, 395]. The imaging area is $\Omega ={{L}_{x}}{{L}_{y}}$ and, since these length scales are not large, one must resort to finite size scaling to compare theory to observation.

Figure 38.

Figure 38. Reprinted from [395] with permission. Copyright 2006 Nature Publishing Group. A sketch of the experimental setup of the interference experiment [395] with coordinate system of text. The imaging area ${{L}_{x}}{{L}_{y}}$ of the individual condensates are the dark shaded areas in the xy planes.

Standard image High-resolution image

In the usual setup, all the physics is contained in the two point correlation $\langle {{a}^{\dagger}}\left(\mathbf{r}\right)a\left(\mathbf{0}\right)\rangle $ which is where the BKT physics appears in the cold atom system. It is well known and verified that the essential behavior in 2D at low T is

Equation (260)

and one sees that,

Equation (261)

Thus, the universal discontinuity of the 2D planar rotor model appears in the context of cold atom physics and is measured in [395] by fitting the density distribution with

Equation (262)

where c(x) is the contrast or amplitude. The average central contrast c0 acts as an effective temperature and is effectively varied by the chosen imaging length Lx [395]. Thus, the physics is encoded in the two measurable parameters ${{c}_{0}},\eta $ which can be varied by an experiment, sketched in figure 38, measuring the integrated contrast

Equation (263)

where $\alpha =0.50$ for ${{c}_{0}}<c_{0}^{\ast}$ or $T>{{T}_{\text{c}}}$ and $\alpha \leqslant 0.25$ for ${{c}_{0}}\geqslant c_{0}^{\ast}$ . The measured values of the exponent $\alpha \left({{c}_{0}}\right)$ are very noisy with large error bars but are certainly consistent with the predictions of KT theory [395]. It is interesting to speculate how well the theory will be able to explain future more detailed and accurate observations on cold atom systems.

6. Summary and acknowledgements

In this review, I have attempted to summarize many areas of 2D physics where Kosterlitz–Thouless theory is relevant to experimental observations and, although I am naturally biased, I am very pleased to discover that the theory still is relevant to several diverse bits of physics and that it can even be quantitively accurate in some situations, which is more than many other theories are! Despite its venerable age of almost half a century, occasionally new physics to which it is applicable appears. Very likely there are applications not mentioned in this review for which I take full responsibility and apologize to those people whose work I either did not mention or have described incorrectly.

I would like to acknowledge my many teachers, collaborators and colleagues especially D Thouless, L Kadanoff, J Reppy, M Fisher, D Fisher, M Fisher, P Anderson, D Mermin, D Nelson, B Halperin, S Doniach, T Ala-Nissila, M Grant, E Granato, J Lee, M Simkin, N Akino, D Obeid and of course my wife who unfailingly supported me during a very long period of neglect, chaos and irritability while this review was being written.

Please wait… references are loading.