Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Two-Dimensional Crystals
Two-Dimensional Crystals
Two-Dimensional Crystals
Ebook695 pages6 hours

Two-Dimensional Crystals

Rating: 0 out of 5 stars

()

Read preview

About this ebook

This is a self-contained, tutorial introduction to two-dimensional crystal science and technology. Including concise descriptions of experimental methods and results from fundamental theoretical concepts, this book covers a broad range of two-dimensional structures--from overlayers to freestanding films. All those with an active interest in surface science and statistical physics will find this book to be an essential reference work.
  • Presents a coherent overview of experimental methods and theoretical background of two-dimensional crystal physics
  • Provides a tutorial overview of continuous melting of two-dimensional crystals, roughening transitions, wetting phenomena, and commensurate-incommensurate transitions
LanguageEnglish
Release dateDec 2, 2012
ISBN9780080924397
Two-Dimensional Crystals

Related to Two-Dimensional Crystals

Related ebooks

Physics For You

View More

Related articles

Reviews for Two-Dimensional Crystals

Rating: 0 out of 5 stars
0 ratings

0 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Two-Dimensional Crystals - A. G. Naumovets

    1

    Preface

    One more book on crystals, among the multitude already published? Yes, it is on crystals, but these are two-dimensional. But are there such crystals? The idea suggests something artificial, perhaps even exotic, deserving little or no attention. Such an assumption is natural, but in fact two-dimensional crystals are omnipresent. They are around us and even on us. Surfaces are their favorite media. Two-dimensional crystals grow on liquids and solids. The reader can infer their frequent occurrence in adsorbed layers from published handbooks of surface structures. They also form from Abrikosov’s vortices in superconducting films and from electrons trapped at a liquid helium surface. This multitude of structures has attracted the attention of the authors of numerous reviews, which are, in part, referred to in the present book. Must we delve into this subject once more? Our belief is that there is room for a new treatment. This book does not treat two-dimensional crystallography as such, but rather the formation of two-dimensional crystals and the phase transitions therein, which are controlled by simple physical mechanisms different from the 3D ones and possessing peculiarities deserving special attention.

    The two-dimensional world has become a vivid reality of our times. Two-dimensional conductors have become the main object of semiconductor technology. Two-dimensional and quasi-two-dimensional magnetics are the focus of attention of specialists designing magnetic memory devices. Almost all high-temperature superconductors incorporate weakly coupled copper oxide planes where the charged carriers form pairs and move. However, two-dimensional crystals are just a small fraction of this world. A complete book on two-dimensional physics would concern itself, for example, with the purely theoretical aspects, from the quantum Hall effect to the conformal field theory and the exactly solved models, which call for the high level of specialization. This abundance of topics requires us to select our goals with a measure of caution. We treat the physics of crystals and digress into the related disciplines only when required to derive fundamental relationships underlying the whole theory of two-dimensional systems.

    Theory and experiment in the field of the two-dimensional crystals developed independently for a long time. This book is an attempt to bridge the gap between them. We have tried to make it readable and interesting for both experimentalists and theoreticians in the field. We hope also that our book will be useful as an introduction to the fascinating world of two-dimensional crystals for nonexperts, e.g., graduate students.

    We have not tried to compile a comprehensive list of references. The choice is inevitably subjective, and we apologize to those authors whose articles are not mentioned.

    This book is based on the Russian edition. It has been difficult to explain in a foreign tongue matters that were not so easy to explain even in Russian. We hope that our efforts have not been in vain.

    This book would never have been produced without support from and cooperation with our friends and colleagues P. Bak, E. Bauer, H. Beck, R. Birgeneau, S. A. Burkov, L. A. Bol’shov, S. Fain, A. G. Fedorus, M. V. Feigel’man, A. Luther, P. Martinoli, V. K. Medvedev, T. Nattermann, M. Schick, Ya. G. Sinai, A. L. Talapov, G. V. Uimin, Yu. S. Vedula, and J. Villain, to whom we express our profound gratitude. Our sincerest thanks go to Greg Dash for his invaluable support in the publication of this book. We are grateful to O. Kiyaev and D. Zateykin for help in the preparation of figures and to N. Nikolayev, I. Tereshko, and N. Konkina for help in the translation of the book. I. F. Lyuksyutov thanks the Alexander von Humboldt Foundation and SFB237 at Ruhr University, and SFB341 at Cologne University, for financial support at last stages of this work.

    Introduction

    Investigations of two-dimensional crystals are traditionally linked to the study of surfaces. The fundamentals of surface thermodynamics were developed by Gibbs as far back as the end of the nineteenth century. Two-dimensional phases and phase transitions between them were first observed by Langmuir in the 1920s in experiments with layers of organic molecules (salts of fatty acids) on the surface of a liquid. These experiments resulted in the indirect discovery of a crystalline phase in a two-dimensional system.

    Major contributions to the understanding of the two-dimensional crystals, from both the theoretical and the experimental viewpoint, date back to the 1920s and 1930s. First, Davisson and Germer succeeded in directly observing the crystalline structure of a surface in their classic experiments on the diffraction of electrons. Second, Landau and Peierls formulated a theorem on the impossibility of the existence of 2D crystals.

    The rapid development of experimental research in the physics of two-dimensional systems began in the 1960s with the emergence of a multitude of methods for diagnostics of surface and other two-dimensional objects, permitting the study of these phenomena on the atomic level. The growth of interest in two-dimensional systems was, to a great extent, stimulated by the requirements of industry (microelectronics, emission electronics, the growth of crystals, catalysis, etc.).

    In the late 1960s, the question of the character of ordering in two-dimensional systems with the continuous group of symmetry was reconsidered by Stanley and Kaplan, Mermin (1968), Hohenberg, Berezinsky (1971a, b), and Jancovici (1967). It was established that the logarithmic divergence of the atomic displacement fluctuations in a 2D crystal that was discovered by Landau and Peierls implies only the absence of long-range order. This divergence manifests itself in the power-law decay of the correlation function of the displacement and, as a result, in the form of the diffraction spots. In spite of this fact, the shear modulus in the 2D crystal does not equal zero.

    The list of 2D crystals includes objects of radically different nature. The longest part of this list consists of the lattices formed by atoms adsorbed on a crystal surface. Two-dimensional structures on surfaces are also formed by atoms of the crystal itself, e.g., during the reconstruction of a surface. Other, less numerous classes of two-dimensional crystals include those formed by electrons trapped at the surface of liquid helium and predicted by Wigner in 1934, though for the 3D case. There are also two-dimensional lattices formed by Abrikosov vortices in superconducting films, colloidal crystals, and crystals in smectic layers and in the above mentioned Langmuir films.

    The variety of these systems means that the characteristic scales of parameter values differ by at least several orders of magnitude for different classes of two-dimensional crystals. Thus, the adatom lattice periods are on the order of several angstroms, whereas the periods of colloidal crystals are on the order of several micrometers. The melting temperature of an electron lattice trapped at a helium surface is a few hundred millikelvin, whereas that of adatom lattices is several hundred kelvins.

    Strong fluctuations in 2D crystals preclude many approaches conventionally used in the study of 3D crystals. The same is true of two-dimensional systems in general. Two-dimensional crystals are a special case of systems with the abelian symmetry group. Therefore, the development of the theory of 2D crystals is closely related to the evolution of two-dimensional mathematical physics. Following the solution of the Ising model by Onsager in 1944, the next major success in this area was the solution of the special case of the six-vertex model, the so-called ice model. This result has stimulated the series of works devoted to the exact solution of the general six-vertex problem, crowned by the solution of the eight-vertex model by Baxter in 1971. This model is referred to as the Baxter model. The works of Berezinsky on the model of plane rotators, also called the XY model, relate to the same period. The ideas of Berezinsky concerning the character of a phase transition in the XY model were addressed in the works by Kosterlitz and Thouless. These ideas were further developed in the works by Jose, Kadanoff, Nelson, and Kirkpatrick (1977) and Wiegman (1978) on the XY and Zn models, and also in the works by Young, Halperin, and Nelson on the melting of an isotropic two-dimensional crystal.

    In the multitude of two-dimensional structures one can distinguish a small, but fundamentally important, class of free two-dimensional crystals. Among these are the freely suspended smectic films of liquid crystals. It is for these films that Moncton et al. (1979, 1982) experimentally proved in the late 1970s the existence of a two-dimensional crystal with a finite shear rigidity characterized, however, by the power-law decay of the correlation function of displacements.

    For the existence of most 2D crystals an external field is needed, to ensure their two-dimensionality. Occasionally, this role is played by the matrix of a 3D crystal containing weakly interacting planes. Intercalated graphite exemplifies such a quasi-two-dimensional system.

    Most 2D crystals are formed on the surface of a liquid or a crystal. Langmuir films and the Wigner crystal exemplify the systems with a liquid substrate. The correlation properties of such systems are identical to those of free two-dimensional crystals, since the surface of a liquid creates no periodic potential corrugation.

    In the case of a crystal substrate the periodic potential corrugation engenders an enormous variety of structures both in adsorbed layers and on clean surfaces. The behavior of these lattices is greatly dependent on the relation of their periods to those of a substrate. For this reason, the whole set of structures found experimentally is divided into two classes: commensurate structures with periods that are multiples of the substrate periods, and incommensurate structures with at least one period incommensurate with the substrate’s period.

    The theory of incommensurate structures goes back to the work by Kontorova and Frenkel in 1938. In 1949, Frank and van der Merwe demonstrated the existence of a lattice of misfit dislocations, or solitons, in a one-dimensional incommensurate chain. Incommensurate structures were discovered experimentally in the 1960s. First came helical magnetic structures, earlier predicted by Dzyaloshinsky, and then incommensurate two-dimensional crystals. Later on, incommensurate three-dimensional crystals were discovered. The physics of incommensurate crystals advanced rapidly in the 1970s and early 1980s, and this development resulted in the discovery of an entirely new type of a structure, i.e., quasicrystals, by Shechtman et al.

    The spectrum of an incommensurate crystal contains an acoustical band representing the displacement of the whole crystal relative to the potential corrugation. Therefore, the two-dimensional crystal is supposed to behave as a free crystal, i.e., the correlation function should show a power-law decay. In 1983, Birgenau and colleagues have demonstrated this fact by investigating x-ray diffraction spot profiles from the incommensurate crystal formed by xenon on the basal plane of graphite. The fundamentals of the statistical mechanics of two-dimensional incommensurate crystals were developed in the works by many authors listed in Chapter 7 together with experimental results.

    The statistical mechanics of commensurate structures was begun by the work of Onsager on the Ising model. At the time of its appearance, Onsager’s theory seemed somewhat rarefied. However, in the late 1960s and early 1970s its experimental realizations were discovered—first among the magnetic systems, and later among the 2D crystals. Physical realizations of many exactly solved models in two-dimensional statistical mechanics (e.g., the Potts, Baxter, and Ashkin–Teller models, the anisotropic Heisenberg model, and the hard-hexagon model) were found among commensurate two-dimensional crystals. This correspondence was established in the works by Alexander (1975) and Domany, Schick, Walker, and Griffiths (1978) and corroborated experimentally.

    Berezinsky (1971a,b) discovered the important role of topological defects in the physics of 2D systems. Later on, Kosterlitz and Thouless (1973) realized that the phase transition in the XY system could be interpreted as a spontaneous creation of vortices. A similar role in the crystal’s melting is played by dislocations. Another type of topological phase transition is related to the spontaneous formation of domain walls. It is a transition from a commensurate to an incommensurate crystal. The spontaneous creation of domain walls also prevails in the phase transition in Ising’s magnetics. The topological defects in systems with a nonabelian symmetry group have a finite energy and cause a transition only at zero temperature.

    We have already noted that experimental research on surfaces and theoretical research on two-dimensional mathematical physics proceeded virtually independently of each other. Therefore, despite our attempts at a unified treatment, the chapters devoted to the description of experimental methods (Chapter 2), surface structures (Chapter 3), and the relation between the structure and the properties of a surface (Chapter 12) are fairly self-contained. For consistency, everywhere in this book we have chosen units such that Boltzmann’s constant kB = 1. Where relevant, the description of an experiment is presented along with the theory.

    The makeup of this book reflects, to a certain extent, the proportions among the number of the experimental investigations of different classes. Free two-dimensional crystals are dealt with in Chapter 4, which gives both the theory of these systems and the standard experimental methods. Chapter 5 describes phases and phase transitions in commensurate crystals. Much more attention is paid to incommensurate crystals (see Chapters 6, 7, and 8). Chapter 9 treats the roughening phase transitions on a crystal’s surface bringing about the changes in its equilibrium shape.

    The effect of various random fields grows with the decline in the dimensionality of a system. The nonequilibrium frozen defects of a substrate change the behavior of a two-dimensional crystal. This subject is treated in Chapter 10. Chapter 11 analyzes the phenomena related to the transition from the 2D to the 3D state, i.e., wetting and surface melting.

    Chapter 1

    Order and Disorder in Two-Dimensional Crystals

    1.1 Two-Dimensional Crystals: Experimental Examples

    To date, many objects of different origin exhibiting a two-dimensional translational order have become known. We refer to these as two-dimensional crystals. Some of them are confined in the third spatial dimension and can be regarded as two-dimensional crystals proper. In contrast, one may have media with a two-dimensional order in the three-dimensional space (Landau and Lifshitz, 1980). The vortex lattices in superfluid ⁴He and bulk superconductors are common examples of the latter. In what follows we shall focus on the two-dimensional crystals proper, omitting the word proper with reference to them.

    The multitude of two-dimensional crystals can be reduced to two unequal classes. The first class includes the free two-dimensional crystals and the two-dimensional crystals whose correlation properties are not influenced by the substrate. An evident example of a free two-dimensional crystal is a film of a smectic liquid crystal, just a few molecules thick, freely suspended in a frame and resembling an ordinary soap film. Less obvious examples are 2D crystals formed by a layer of electrons trapped at a height of the order of 100 Å above the surface of liquid helium, and Abrikosov vortex lattices in superconducting films.

    The 2D crystals whose correlation properties are affected by the substrate (a three-dimensional base) form the second class and are much more numerous than free crystals. In many systems, the atoms adsorbed on the surface of a crystalline substrate actually form ordered overlayers at early stages of adsorption. Such adatom lattices in overlayers form the most numerous and the best-studied class of 2D crystals (see, e.g., Ohtani et al., 1986). The two-dimensionality of these systems results from the penetration of atom-caused deformation into the substrate to the depth of a few lattice constants. The structure of surface layers of some clean adsorbateless metals and semiconductors is different from that of layers in the bulk of the crystal (in this case the surface is said to undergo reconstruction). Surface layers of this kind also exemplify two-dimensional crystals on a substrate. As in the overlayers, their two-dimensionality is caused by the shallow intrusion of surface-reconstruction-induced deformations into the bulk of a crystal to the depth of just a few lattice constants.

    A number of systems which can be regarded as quasi-two-dimensional ones have been found experimentally. They can be regarded as stacks of 2D crystals, and their behavior exhibits properties characteristic of two-dimensional crystals. The reason for this similarity is the weakness of the interaction between the layers. Graphite intercalated by different atoms is a typical example of a quasi-two-dimensional crystal.

    Let us now characterize in greater detail the systems treated in this chapter.

    1.1.1 Freestanding Smectic Liquid Crystal Films

    Films of this kind are like ordinary soap films. Under real experimental conditions they are suspended in frames, have a free surface of the order of 1 cm², and are of virtually uniform thickness. The films are composed of layers of rodlike molecules oriented normal to the film’s plane (Fig. 1.1). The least number of layers attained experimentally equals two, with a total thickness of about 50 Å. In experiments with a controlled number of layers this figure varied from two to a few hundred. Molecules of the liquid crystal form a hexagonal lattice in the film plane, which melts as the temperature increases. The structural, thermodynamic, mechanical, and correlation properties of such lattices have been studied experimentally (see Sections 1.3 and 4.3). If the number of layers is small, the films exhibit properties characteristic of two-dimensional crystals.

    Figure 1.1 Sketch of experimental setup for formation of smectic films a few molecular layers thick: 1, liquid crystal; 2, spreader for application of the film; 3, spreads motion direction; 4, window with the film of liquid crystal ( Brinkman et al. , 1982 .) (Copyright 1982 by the AAAS.)

    1.1.2 Lattices of Adatoms Adsorbed on Metal Surfaces

    There exists a fairly numerous family of two-dimensional crystals formed on metals. Let us consider a typical example—the La − W(112) system (Vedula et al.direction (Fig. 1.2a). In this system lanthanum forms a series of structures: p(l × 7) (Fig. 1.2b), c(2 × 2) (Fig. 1.2c), usually referred to as commensurate structures, whose periods are multiples of periods of the substrate’s crystal. The letters p and c denote the primitive and the centered elementary cells, respectively, while the numbers indicate the ratios of periods of the surface and substrate crystals. As the coverage in the overlayer increases, the interaction between adatoms becomes stronger and they form a so-called incommensurate lattice (Fig. 1.2d) whose period along the furrows varies continuously with coverage being, generally, an irrational multiple of a substrate’s period. Such a mathematical definition of the incommensurate crystal might seem not very clear to a physicist, who measures any quantity with only a certain accuracy. Its physical meaning is that parameters of the lattice of the two-dimensional crystal can be continuously varied with respect to the fixed, constant parameters of the substrate lattice. The structure of the two-dimensional crystals formed on metal surfaces is usually studied by the low-energy electron diffraction (LEED) technique.

    Figure 1.2 Overlayer structures in the system La − W(112): a, clean surface; b, p ); c, c ), d, incommensurate structure. ( Vedula et al. , 1977 b .)

    1.1.3 Overlayers on Exfoliated Graphite

    By a special processing method graphite is cleaved along the basal plane into thin plates with surfaces of a high degree of perfection, so that uniform areas are sized a few thousand angstroms. Exfoliated graphite can have an enormous specific surface, tens of m²/g, which allows effective employment of x-ray diffraction for studies of the two-dimensional crystals formed by atoms of various adsorbates on the basal plane of graphite (see Section 2.1). The enormous specific surface of the exfoliated graphite also permits calorimetric studies of two-dimensional crystals. Finally, using monocrystal-line graphite substrates, one can study the two-dimensional crystals by the LEED technique. As a rule, the adsorbates used are gases, mostly noble ones, whose interaction with the graphite substrate is weak. An example of the incommensurate lattice formed by xenon adatoms on the basal plane of graphite (Heiney et al., 1983) is shown in Fig. 1.3. This lattice, like the substrate, has a hexagonal symmetry.

    Figure 1.3 Hexagonal incommensurate lattice of xenon atoms on the basal plane of graphite (black and open circles correspond to graphite and xenon atoms, respectively).

    1.1.4 Intercalated Graphite

    In this system, the two-dimensional crystals are formed by layers of atoms implanted (intercalated) into graphite. These layers are arranged between the adjacent basal planes of the graphite. A spacing of layers may equal a few graphite lattice constants, so that interaction between atoms of different layers is weak compared to that between atoms of the same layer, thus providing the quasi-two-dimensionality of the system. Shown in Fig. 1.4 is the location of bromine atoms in the unit cell of the two-dimensional crystal formed in the compound C28Br2 (Erbil et al., 1983). They form a commensurate lattice with the period in direction OY equal to 7 periods of the graphite lattice. The spacing of bromine layers is four periods of the graphite lattice. The bromine concentration is changed by varying either the pressure of bromine vapor in the sample chamber or the temperature. For instance, when the pressure is raised, the bromine lattice becomes incommensurate along the OY axis. Since intercalated compounds are stacks of two-dimensional crystals, their structure can be studied by x-ray diffraction methods.

    Figure 1.4 Elementary cell of the commensurate structure c (√3 × 7) of bromine atoms intercalated in graphite. ( Erbil et al. , 1983 .)

    We have listed some cases of two-dimensional crystals. This list can easily be expanded to include the lattices of Abrikosov’s vortices in superconducting films, the Langmuir films, interfaces, grain boundaries, etc. Though of different origin, these systems belong to the class of two-dimensional crystals in that they share one principal signature—correlation properties. Whether or not the system should be regarded as two-dimensional (or quasi-two-dimensional) may depend on which features are considered important. For instance, the system may well be two-dimensional with regard to its electrical conductivity or magnetic properties, but it is three-dimensional with regard to elasticity. The meaning of the two-dimensional signature of the correlation properties, and the mechanisms whereby they are revealed in the experimentally measured properties, are discussed in subsequent sections of this chapter.

    1.2 Structure of Two-Dimensional Crystals and Substrate

    A substrate is an integral part of a majority of two-dimensional crystals. Its presence engenders a great variety of structures of two-dimensional crystals, as they depend both on the interaction between atoms of the crystal itself and on the interaction with the substrate atoms. Overlayers are notable in this respect. As a rule, a substance in an adsorbed state possesses many more lattices of different symmetry than does a three-dimensional substance. Two-dimensional crystalline phases are observed over a very broad range of densities. Thus, periods of two-dimensional adsorbate lattices can vary from 25 Å (the largest experimentally observed period), to values below the period of the corresponding three-dimensional crystals.

    In this section we shall discuss general ideas on the influence of the substrate on the structure of two-dimensional crystals. Though for the sake of definiteness we consider adatom lattices, no generality is lost. As the interaction between adatoms is usually much weaker than that between the substrate atoms, it can be assumed that the substrate produces a rigid periodic potential corrugation or potential relief for adatoms. One can single out two rough but nonetheless well-defined parameters that are measures of the effect of the substrate potential corrugation on the structure of a two-dimensional crystal. The first one is a ratio of the mean energy of interaction of adatoms, J, to the substrate potential corrugation amplitude V. The second one is the coverage θ, i.e., the ratio of the number of atoms in the crystal to the number of minima of the substrate potential. These two parameters are not independent, since the higher θ is, the smaller the interatomic distance and the larger J will be. The two limiting cases of weak (J V) and strong (J V) substrate potential corrugation have been studied most extensively. In the first case the potential can be treated as a perturbation of the free two-dimensional crystal. In the second case adatoms of the crystal occupy exactly the minima of the substrate potential, which also simplifies the analysis. Quite often both limiting cases are attainable in the same system, since the two-dimensional crystals exist over a wide range of θ, as has been mentioned.

    The simplest situation occurs when the vectors of the unit cell of a crystal are multiples of vectors of the substrate lattice, so that all the adatoms occupy minima of the potential. This takes place at certain special values of the coverage θ. For instance, in the system La − W(112) (see (Fig. 1.2c). A natural question arises: what does the structure look like at an arbitrary θ? According to the Gibbs rule (Landau and Lifshitz, 1980), with the two thermodynamic variables—temperature T and coverage θthe lattice gas and crystalline phase p(1 × 7) coexist (naturally at temperatures below the melting point of the structure). The single-phase system at an intermediate θ (Fig. 1.2d) the compression of the structure along the furrows is observed. This implies that interaction between adatoms starts to exceed the substrate potential corrugation. As a result, an incommensurate crystal forms. Let us consider in greater detail how this process occurs.

    1.2.1 Incommensurate Two-Dimensional Crystal

    It is natural to consider this problem for weak substrate potential corrugation. For the sake of clarity, the adatoms are supposed to form a rectangular lattice with a primitive unit cell contracted along the OX axis (see Fig. 1.5). This is a model of a uniaxial incommensurate crystal, which has a number of experimental realizations. In this model, the atoms are supposed to shift only along the OX direction. Let u be a displacement. In the continuum medium approximation, the elastic deformation energy is of the form

       (1.1)

    Figure 1.5 Positions of atoms in a soliton formed in a uniaxial incommensurate crystal (dashed circles). At the bottom u is plotted against x .

    The first term in (1.1) describes the compression of the lattice along the OX axis, whereas the second one describes its bending. The corresponding elasticity constants are in general unequal, but they can always be rendered identical by proper choice of scales along the OX and OY axes.

    Let us consider now the crystal − substrate interaction in the rigid substrate approximation. In this case, the net effect of the substrate is a periodic potential for adatoms. We shall use this approximation extensively, leaving to Section 6.8 the consideration of effects of elastic deformations of the substrate. For a uniaxial crystal, the energy of interaction with the substrate will vary periodically with the x-coordinate only, the period being that of the substrate. The adatom-lattice − substrate interaction energy has a form most convenient for further analysis in the case of the cosine substrate potential corrugation:

       (1.2)

    Here a and b are the periods of the crystal and substrate along the OX axis, and V is an amplitude of the substrate potential corrugation. We shall suppose nearly equal periods a and b ((a − b)/b 1). A more general case where the ratio a/b is close to an arbitrary rational number will be studied in Chapter 6. The argument of the cosine in (1.2) has a simple origin. The first term comes from the regular displacement of atoms of an undeformed crystal lattice relative to the substrate potential, equal to (b − a)x/a at a distance x. The second term is the displacement u(x) of atom at point x because of the lattice deformation, divided by the lattice period.

    The ground state of the two-dimensional crystal is determined from the condition of minimal total energy

    Standard variations of H in u(x, y) lead to the equation for the equilibrium state,

       (1.3)

    Before the formal analysis of the solution of equation (1.3), which describes the incommensurate phase, some qualitative considerations are in order. Enlarging the coverage θ results in more atomic rows. If initially a = b, i.e., θ = 1, then at θ > 1 adatoms are displaced from the minima of the substrate potential. However, it is fairly evident that after adding one extra row it is inexpedient to displace all the atoms from the minimal energy positions. It would be more favorable if the deformed area were localized in a strip along the OY axis with a finite width lo along the OX axis. Then the deformation and deformation energy per lattice period along OY x ~ b/lo and

    On the other hand, the increase of potential energy of interaction with the substrate is of the order of Vl0. Minimization of the total energy yields

    Therefore, it is energetically favorable for deformations caused by an excess row of adatoms to be localized in a strip of width l0. Such a strip is called soliton or domain wall. Outside the soliton, atoms are located at the minima of the substrate potential, so that the soliton separates domains of the commensurate phase.

    The exact one-soliton solution found from (1.3) is

       (1.4)

    and fully corroborates the above qualitative considerations. Since the substrate potential depends only on x, the problem of the ground state of the uniaxial incommensurate crystal is close to the problem of one-dimensional incommensurate chain. The latter problem was first considered by Kontorova and Frenkel (1938) and later on by Frank and van der Merwe (1949a, b). If more than one excess row is added, then evidently a periodic lattice of solitons will be produced. If the coverage in the experiment is fixed, the period l of the soliton lattice will be a function of θ. In the above example l = b/(θ − 1). As θ rises, the lattice period will decrease until the very concept of soliton will lose its sense at l ~ l0. In this limit the incommensurate lattice is practically homogeneous.

    In the incommensurate phase there appears a novel, quasiacoustical branch in the excitation spectrum—a manifestation of the absence of a barrier for translation of the soliton lattice. This can be easily checked by substituting the solution (1.4) for an isolated soliton into the formula for the energy (1.1). The energy of the soliton is independent of its position on the substrate. This can easily be understood qualitatively for the absolutely rigid incommensurate crystal. Such a lattice will be periodic in a strict sense, so that in view of incommensurability, the potential energy of atoms takes all the values in an interval from − V to V. For this reason translations of the lattice through any distance cause a redistribution of energy between atoms, rather than a change in the total energy, i.e., the substrate is effectively flat for the incommensurate

    Enjoying the preview?
    Page 1 of 1