Self-organization of adatom adsorption structure at interaction with tip of dynamic force microscope

The formation of an adatom adsorption structure in dynamic force microscopy experiment is shown as a result of the spontaneous appearance of shear strain caused by external supercritical heating. This transition is described by the Kelvin-Voigt equation for a viscoelastic medium, the relaxation Land...

Ausführliche Beschreibung

Gespeichert in:
Bibliographische Detailangaben
Datum:2014
1. Verfasser: Khomenko, A.V.
Format: Artikel
Sprache:English
Veröffentlicht: Інститут фізики конденсованих систем НАН України 2014
Schriftenreihe:Condensed Matter Physics
Online Zugang:http://dspace.nbuv.gov.ua/handle/123456789/152897
Tags: Tag hinzufügen
Keine Tags, Fügen Sie den ersten Tag hinzu!
Назва журналу:Digital Library of Periodicals of National Academy of Sciences of Ukraine
Zitieren:Self-organization of adatom adsorption structure at interaction with tip of dynamic force microscope / A.V. Khomenko // Condensed Matter Physics. — 2014. — Т. 17, № 3. — С. 33401: 1–10 — Бібліогр.: 50 назв. — англ.

Institution

Digital Library of Periodicals of National Academy of Sciences of Ukraine
id irk-123456789-152897
record_format dspace
spelling irk-123456789-1528972019-06-14T01:25:28Z Self-organization of adatom adsorption structure at interaction with tip of dynamic force microscope Khomenko, A.V. The formation of an adatom adsorption structure in dynamic force microscopy experiment is shown as a result of the spontaneous appearance of shear strain caused by external supercritical heating. This transition is described by the Kelvin-Voigt equation for a viscoelastic medium, the relaxation Landau-Khalatnikov equation for shear stress, and the relaxation equation for temperature. It is shown that these equations formally coincide with the synergetic Lorenz system, where the shear strain acts as the order parameter, the conjugate field is reduced to the stress, and the temperature is the control parameter. Within the adiabatic approximation, the steady-state values of these quantities are found. Taking into account the sample shear modulus vs strain dependence, the formation of the adatom adsorption configuration is described as the first-order transition. The critical temperature of the tip linearly increases with the growth of the effective value of the sample shear modulus and decreases with the growth of its typical value. Формування структури адсорбованих адатомiв при дослiдженнi в режимi динамiчної силової мiкроскопiї представлено як результат спонтанної появи зсувної деформацiї в результатi зовнiшнього надкритичного нагрiвання. Цей перехiд описується рiвнянням Кельвiна-Фойгта для в’язкопружного середовища, релаксацiйним рiвнянням Ландау-Халатнiкова для зсувних напружень та релаксацiйним рiвнянням для температури. Показано, що цi рiвняння формально збiгаються iз синергетичною системою Лоренца, де зсувна деформацiя вiдiграє роль параметра порядку, спряжене поле зводиться до напружень, та температура є керувальним параметром. В рамках адiабатичного наближення знайденi стацiонарнi значення цих величин. Враховуючи залежнiсть модуля зсуву зразка вiд деформацiї, формування конфiгурацiї адсорбованих адатомiв описано як перехiд першого роду. Критична температура зонда лiнiйно зростає з ростом ефективного значення модуля зсуву зразка i зменшується при зростаннi його характерного значення. 2014 Article Self-organization of adatom adsorption structure at interaction with tip of dynamic force microscope / A.V. Khomenko // Condensed Matter Physics. — 2014. — Т. 17, № 3. — С. 33401: 1–10 — Бібліогр.: 50 назв. — англ. 1607-324X DOI:10.5488/CMP.17.33401 PACS: 46.55.+d, 64.60.-i, 61.72.Hh, 62.20.F-, 62.20.Qp, 68.37.Ps arXiv:1301.4379 http://dspace.nbuv.gov.ua/handle/123456789/152897 en Condensed Matter Physics Інститут фізики конденсованих систем НАН України
institution Digital Library of Periodicals of National Academy of Sciences of Ukraine
collection DSpace DC
language English
description The formation of an adatom adsorption structure in dynamic force microscopy experiment is shown as a result of the spontaneous appearance of shear strain caused by external supercritical heating. This transition is described by the Kelvin-Voigt equation for a viscoelastic medium, the relaxation Landau-Khalatnikov equation for shear stress, and the relaxation equation for temperature. It is shown that these equations formally coincide with the synergetic Lorenz system, where the shear strain acts as the order parameter, the conjugate field is reduced to the stress, and the temperature is the control parameter. Within the adiabatic approximation, the steady-state values of these quantities are found. Taking into account the sample shear modulus vs strain dependence, the formation of the adatom adsorption configuration is described as the first-order transition. The critical temperature of the tip linearly increases with the growth of the effective value of the sample shear modulus and decreases with the growth of its typical value.
format Article
author Khomenko, A.V.
spellingShingle Khomenko, A.V.
Self-organization of adatom adsorption structure at interaction with tip of dynamic force microscope
Condensed Matter Physics
author_facet Khomenko, A.V.
author_sort Khomenko, A.V.
title Self-organization of adatom adsorption structure at interaction with tip of dynamic force microscope
title_short Self-organization of adatom adsorption structure at interaction with tip of dynamic force microscope
title_full Self-organization of adatom adsorption structure at interaction with tip of dynamic force microscope
title_fullStr Self-organization of adatom adsorption structure at interaction with tip of dynamic force microscope
title_full_unstemmed Self-organization of adatom adsorption structure at interaction with tip of dynamic force microscope
title_sort self-organization of adatom adsorption structure at interaction with tip of dynamic force microscope
publisher Інститут фізики конденсованих систем НАН України
publishDate 2014
url http://dspace.nbuv.gov.ua/handle/123456789/152897
citation_txt Self-organization of adatom adsorption structure at interaction with tip of dynamic force microscope / A.V. Khomenko // Condensed Matter Physics. — 2014. — Т. 17, № 3. — С. 33401: 1–10 — Бібліогр.: 50 назв. — англ.
series Condensed Matter Physics
work_keys_str_mv AT khomenkoav selforganizationofadatomadsorptionstructureatinteractionwithtipofdynamicforcemicroscope
first_indexed 2025-07-14T04:21:53Z
last_indexed 2025-07-14T04:21:53Z
_version_ 1837594747057733632
fulltext Condensed Matter Physics, 2014, Vol. 17, No 3, 33401: 1–10 DOI: 10.5488/CMP.17.33401 http://www.icmp.lviv.ua/journal Self-organization of adatom adsorption structure at interaction with tip of dynamic force microscope A.V. Khomenko1,2 1 Department of Complex Systems Modelling, Sumy State University, 2 Rimskii-Korsakov St., 40007 Sumy, Ukraine 2 Peter Grünberg Institut-1, Forschungszentrum-Jülich, 52425 Jülich, Germany Received April 5, 2014, in final form June 04, 2014 The formation of an adatom adsorption structure in dynamic force microscopy experiment is shown as a re- sult of the spontaneous appearance of shear strain caused by external supercritical heating. This transition is described by the Kelvin-Voigt equation for a viscoelastic medium, the relaxation Landau-Khalatnikov equation for shear stress, and the relaxation equation for temperature. It is shown that these equations formally coin- cide with the synergetic Lorenz system, where the shear strain acts as the order parameter, the conjugate field is reduced to the stress, and the temperature is the control parameter. Within the adiabatic approximation, the steady-state values of these quantities are found. Taking into account the sample shear modulus vs strain dependence, the formation of the adatom adsorption configuration is described as the first-order transition. The critical temperature of the tip linearly increases with the growth of the effective value of the sample shear modulus and decreases with the growth of its typical value. Key words: phase transition, rheology, plasticity, strain, stress, atomic force microscopy PACS: 46.55.+d, 64.60.-i, 61.72.Hh, 62.20.F-, 62.20.Qp, 68.37.Ps 1. Introduction Nowadays, due to large scientific and practical importance, the phenomena taking place on the sam- ple surface at interaction with the tip of a dynamic force microscope, e.g., atomic force microscope (AFM) and friction force microscope, attract more andmore attention (see the reviews in [1–6] and the literature cited therein). Particularly, the experimental and theoretical data are obtained on structural instabilities, phase transformations, plastic dislocation, neck and adatom structures formation [7–20]. These processes are characterized by hysteresis of dependencies of adhesion force and potential energy surface on the tip- surface distance [11, 16, 21–24] and by hysteresis of the sample stress vs strain curve [25]. Since the nature of such phenomena remains poorly understood, the basic goal of the present study is the construction of a qualitative nonlinear model [26–31] describing the hysteresis processes which occur on the germanium surface during interaction with the AFM tip [11]. Here, the macroscopic con- tinuum mechanics models [32, 33] are supposed to be still applicable to the atomic length-scales, where discrete atomistic interactions become significant [3, 16, 18, 34, 35]. However, a total explanation of the studied macroscopic phenomena requires a consideration of microscopic processes. Phenomenological description used here makes it possible to connect the parameters of microscopic theories with macro- scopic measurements. However, this is a separate independent problem that is hard to solve just now. In the presented approach, the formation conditions of the adatom adsorption structure are defined on the semiconductor surface due to both thermal and deformation effects. The total set of freedom degrees is considered as equivalent variables. The adatom configuration formation is described analytically as a result of self-organization caused by the positive feedback of shear strain and temperature on shear stress on the one hand, as well as the negative feedback of shear strain and stress on temperature on the © A.V. Khomenko, 2014 33401-1 http://dx.doi.org/10.5488/CMP.17.33401 http://www.icmp.lviv.ua/journal A.V. Khomenko other hand. This study is based on the assumption that stress relaxation time diverges because the shear modulus vanishes at the point of transition. The paper is organized as follows. In section 2 the self-consistent Lorenz system of the governed equations is written for approximation of semiconductor characterized by heat conductivity. The adatom structure formation is shown in section 3 to be supercritical in character (is of the second order) when the effective shear modulus of the germanium does not depend on the strain value; it then transforms to a subcritical mode with this dependence appearance (section 4). In these sections, the steady-state values of shear strain and stress, as well as temperature are also determined within adiabatic approximation. Using such a limit, a synergetic potential is obtained, that is the analog of a thermodynamic potential, from basic evolution equations. Section 5 contains short conclusions. 2. Basic equations Let us start with the supposition that the relaxation behavior of the shear component ε of the strain tensor in a semiconductor is governed by the Kelvin-Voigt equation [32, 36] ε̇=−ε/τε+σ/ηε , (1) where τε is the Debye relaxation time and ηε is the effective shear viscosity coefficient. The second term on the right-hand side describes the flow of a viscous liquid caused by the corresponding shear compo- nent of the stress σ. In the steady state, ε̇= 0, we obtain the Hooke-type expression σ=Gεε, Gε ≡ ηε/τε. The next assumption of our approach is that the relaxation equation of the sample shear stress σ has a form similar to the Landau-Khalatnikov equation [28, 30, 37]: τσσ̇=−σ+G(T )ε. (2) Here, the first term on the right-hand side describes the relaxation during time τσ ≡ η/G(T ) determined by the values of the shear viscosity η and modulus G(T ) depending on the sample temperature. In the stationary case σ̇= 0, the kinetic equation (2) is transformed into the Hooke’s law σ=G(T )ε. (3) Note that effective values of viscosity ηε ≡ τεGε and modulus Gε ≡ ηε/τε do not coincide with the real values η and G(T ). Physically, such difference is conditioned by the Landau-Khalatnikov-type equation (2) being not equivalent to the Kelvin-Voigt equation (1) [30, 32, 33]. As is known, the values Gε, η, ηε very weakly depend on the sample temperature T , while the real shear modulus G(T ) vanishes, when the temperature decreases to Tc [38–42]. Further, the simplest approximate temperature dependencies are used: Gε(T ), η(T ),ηε(T ) = const, G(T ) =G0 (T /Tc −1) , (4) where G0 ≡G(T = 2Tc) is the typical value of modulus. According to the synergetic concept [26, 29, 30, 40, 41, 43, 44] to complete the equation system (1) and (2), which contains the order parameter ε, the conjugate field σ, and the control parameter T , we should deduce a kinetic equation for the temperature. This equation can be obtained using the basic relationships of elasticity theory stated in § 32 in [33]. Thus, it is necessary to start with the continuity equation for the heat δQ = TδS: T Ṡ =−∇q. (5) Here, the heat current is given by the Onsager equation q =−κ∇T, (6) where κ is the heat conductivity. In the elementary case of the thermoelastic stress, the entropy S = S0(T )+Kαε0 (7) 33401-2 Self-organization of adatom adsorption structure at interaction with the tip of dynamic force microscope consists of the purely thermodynamic component S0 and the dilatation: ε̂0 = ε0 Î , ε0 ≡α(T −T0) , (8) where α is the thermal expansion coefficient, T0 is the equilibrium temperature, Î is the unit tensor and K is the compression modulus (see § 6 in [33]). In the considered situation, we should transfer from the dilatational component Kαε0 to the elastic energy −σε/T of the shear component divided by tempera- ture (here, the minus sign takes into account the connection TδS = pδV ⇒−σδε at S0 = const, which is caused by the opposite choice of the pressure p and the stress σ signs). As a result, equation (5) has the form T Ṡ0(T )−σε̇=κ∇2T. (9) Taking into account the approximation (κ/l 2)(Te−T ) ≈κ∇2T (l is the scale of heat conductivity, Te is the AFM tip temperature) and the definition of heat capacity cp=T dS0/dT , equation (9) assumes the form: cpṪ = κ l 2 (Te −T )+σε̇. (10) Substituting the expression for the ε̇ from equation (1) we obtain the term σ2/ηε. It describes the dis- sipative heating of a viscous liquid flowing under the effect of the stress σ that can be neglected in the case under consideration. On the other hand, the process of an AFM tip moving into contact with the surface has the following peculiarity. It is necessary to consider the thermal effect of the tip whose value Te is not reduced to the Onsager component and is fixed by external conditions. In view of these circum- stances, the square contribution of the stress is supposed to be included in Te. The obvious account of this term leads to a significant complication of the subsequent analysis, though it results in a renormalization of the quantities. Therefore, component Te in equation (10) is assumed to be constant for our further consideration. It is convenient to introduce the following measure units: σs = ( cpηεTc/τT )1/2 , εs =σs /Gε , Tc , (11) for the variables σ, ε, T , respectively (τT ≡ l 2cp/κ is the time of heat conductivity). Then, the basic equa- tions (1), (2), and (10) take the form: τεε̇=−ε+σ, (12) τσσ̇=−σ+ g (T −1)ε, (13) τTṪ = (Te −T )−σε, (14) where the constant g = G0 Gε (15) is introduced. Equations (12)–(14) have a form similar to the Lorenz scheme [26] which allows us to describe the thermodynamic phase and the kinetic transitions [29, 30, 40, 41, 43–45]. 3. Continuous transition In general the system (12)–(14) cannot be solved analytically. Therefore, we use the following adia- batic approximation: τσ ≪ τε , τT ≪ τε . (16) This implies that in the course of the matter evolution, the stress σ(t) and the temperature T (t) follow the variation of the strain ε(t). The first of these inequalities is fulfilled because it contains the macroscopic time τε and the microscopic Debye time τσ ≈ a/c∼10−12 s, where a ∼ 1 nm is the lattice constant or the 33401-3 A.V. Khomenko intermolecular distance and c ∼ 103 m/s is the sound velocity. The second condition (16) can be reduced to the form l ≪ L, (17) where the maximal value of the characteristic length of the heat conductivity L = √ χνε c2 ε , (18) the thermometric conductivity χ ≡ κ/cp, the effective kinematic viscosity νε ≡ ηε/ρ and the sound ve- locity cε ≡ (Gε/ρ)1/2 are introduced (ρ is the medium density). Then, we can put the left-hand sides of equations (13) and (14) to be equal to zero. As a result, the stress σ and the temperature T are expressed in terms of the strain ε: σ= gε(Te −1) 1+ gε2 , (19) T = 1+ Te −1 1+ gε2 . (20) In accordance with equation (20), in the important range of values of the parameter Te > 1, the temper- ature T decreases monotonously with an increasing strain ε from the value Te at ε = 0 to (Te +1)/2 at ε = εm ≡ √ 1/g . Obviously, this decrease is caused by the negative feedback of the stress and the strain on the temperature in equation (14), which is explained by the Le Chatelier principle for this problem. Really, the reason for the formation of adatom adsorption structure is the positive feedback of the strain and the temperature on the stress in equation (13). Hence, the increase in the temperature should in- tensify the self-organization effect. However, according to equation (14), the system behaves in such a way that the consequence of transition, i.e., the growth of the strain, leads to a decrease in its cause (i.e., temperature). Equation (19), expressing the stress in terms of the strain, has a linear form of the Hooke’s law at ε≪ εm with the effective shear modulus Gef ≡ g (Te −1). At ε= εm , the function σ(ε) has a maxi- mum and at ε> εm it decreases, which has no physical meaning. Thus, the constant εm ≡ √ 1/g gives the maximal strain. An increase in the typical value of the modulus G0 leads to a decrease in the maximal strain εm and to an increase in the effective modulus Gef whose value is proportional to the characteristic temperature Te. Substituting equation (19) into equation (12), we obtain the Landau-Khalatnikov-type equation [28, 37, 46, 47] τεε̇=−∂V /∂ε, (21) where the synergetic potential has the form V = 1 2 [ ε2 + (1−Te) ln ( 1+ gε2 )] . (22) At a steady state, the condition ε̇ = 0 is fulfilled and the potential (22) acquires a minimum. When the temperature Te becomes smaller than the critical value Tc0 = 1+ g−1; g ≡G0/Gε < 1, Gε ≡ ηε/τε , (23) this minimum corresponds to ε = 0, i.e., the adatom adsorption structure is not realized. In the reverse case Te > Tc0, the stationary shear strain has the nonzero value ε0 = [ Te − (1+ g−1) ]1/2 (24) which increases with Te growth in accordance with the root law. This causes the formation of the adatom configuration. Equations (19) and (20) give the stationary values of stress and temperature: σ0 = ε0 , T0 = 1+ g−1. (25) Note that, on the one hand, the steady temperature T0 coincides with the critical value (23) and, on the other hand, its value differs from the temperature Te. Since Tc0 is the minimal temperature at which 33401-4 Self-organization of adatom adsorption structure at interaction with the tip of dynamic force microscope the formation of the adatom adsorption structures can be observed, the above implies that the negative feedback of the stress σ and the strain ε on the temperature T [see the last term on the right-hand side of equation (14)] decreases the sample temperature so much that only in the limit does it ensure the self-organization process. At a steady state, the value of the shear modulus is Gs =Gε . (26) The two cases can be marked out by the parameter g = G0/Gε. In the situation g ≫ 1, meeting the large value of the modulus G0, equations (23)–(25) take the form ε0 = (Te −1)1/2, T0 = Tc0 = 1. (27) This corresponds to the “solid (fragile)” limit. The opposite case g ≪ 1 (small modulus G0) meets the “strongly viscous liquid” ε0 = ( Te − g−1 )1/2 , T0 = Tc0 = g−1 =Gε/G0 . (28) 4. Effect of deformational defect of modulus The Kelvin-Voigt equation (1) assumes the use of the idealized Genki model. For the dependence σ(ε) of the stress on the strain, this model is described by the Hooke’s expression σ=Gεε at ε< εm and by the constant σm = Gεεm at εÊ εm (σm, εm are the maximal stress and strain, σ> σm results in viscous flow with the deformation rate ε̇ = (σ−σm)/ηε). Actually, the σ vs ε dependence curve has two regions: the first one, Hookean, has a large slope corresponding to the shear modulus Gε, followed by a more gently sloping section of the plastic deformation whose tilt is defined by the hardening factor Θ<Gε. Obviously, such a picture means that the shear modulus, introduced in equation (1), depends on the strain value. Let us use the simplest approximation [48, 49] Gε(ε) =Θ+ Gε−Θ 1+ ( ε/εp )2 , (29) which describes the above mentioned transition of the elastic deformation mode to the plastic one. It takes place at a characteristic value of the strain εp, which is smaller than εs (otherwise plastic mode is not realized). Note that an expression of the type equation (29) was originally proposed by Haken [26] describing the rigid mode of laser radiation. It is used [29, 43, 44] to describe the first-order phase transition, and equation (29) contained the square of the ratio ε/εp (so, the V vs ε dependencies in [29, 43, 44] and equation (30) have an even form). In the description of structural phase transitions of a liquid, the third-order invariants, breaking the specified parity, are present [27]. Therefore, in the study [30, 40, 41], in the approximation (29), we used the linear term ε/εp, instead of the square term ( ε/εp )2 . Obviously, in this case, the V vs ε dependence is already uneven. Within the adiabatic approximation (16), the Lorenz equations (12)–(14), where Gε is replaced by a dependence Gε(ε), is reduced to the Landau-Khalatnikov equation (21). The synergetic potential has the form: V = 1 2 ε2 − gα2(Te −1) 2 { 1 gα2 −θ ln ∣∣∣∣ 1+ gε2 1+θ(ε/α)2 ∣∣∣∣ + 1 θgα2(θ−1 − g−1α−2) [ θ−1 ln ∣∣θ−1 + (ε/α)2 ∣∣− g−1α−2 ln ∣∣g−1α−2 + (ε/α)2 ∣∣] } . (30) Here, the constant α≡εp/εs < 1 and the parameter θ=Θ/Gε < 1, describing the ratio of the tilts for the de- formation curve on the plastic and the Hookean sections, are introduced. At a small value of temperature Te , the dependence (30) has a monotonously increasing shape with its minimum at ε= 0 corresponding to the steady state of the absence of the adatom adsorption structure (curve 1 in figure 1). As shown in figure 1, at T 0 c =1+θg−1 +α2(1−2θ)+2α √ g−1θ(1−θ)(1−gα2), (31) 33401-5 A.V. Khomenko Figure 1. Dependence of the synergetic potential on the strain at g=0.2, θ=α=0.25 and various tempera- tures: (curve 1) Te<T 0 c , (curve 2) Te=T 0 c , (curve 3) T 0 c <Te<Tc , and (curve 4) TeÊTc0. Figure 2. Dependence of the steady-state values of the strain on the temperature Te at parameters of figure 1 (the solid curve corresponds to the stable value ε0, the dashed curvemeets the unstable one, εm). a plateau appears (curve 2), which for Te > T 0 c is transformed into a minimum, meeting the strain ε0 , 0, and a maximum at εm that separates the minima corresponding to the values ε= 0 and ε = ε0 (curve 3) [11]. With a further increase in the temperature Te, the “ordered” phase minimum, corresponding to the adatom adsorption configuration ε= ε0, grows deeper, and the height of the interphase barrier decreases vanishing at the critical value Tc0 = 1+ g−1 (23). The steady-state values of the strain have the form (see figures 1 and 2) ( εm 0 )2 = ( 2gθ )−1 { g ( Te −1−α2 ) −θ∓ √[ g ( Te−1−α2 ) −θ ]2 −4gα2θ [ 1−g (Te−1) ]} , (32) where the lower sign meets the stable adatom structure and the upper sign corresponds to the unstable one. At Te Ê Tc0, the dependence V (ε) is characteristic of the absence of the modulus defect (see curve 4 in figure 1). It is worth noting that the potential barrier inherent in the synergetic first-order transition manifests itself only due to the deformational defect of the modulus. Since the latter is realized always [11], it follows that the studied adatom structure formation is a synergetic first-order transition. The considered situation differs from typical thermodynamic phase transitions. Really, in the latter case, the stationary value of the semiconductor temperature T0 is equal to the thermostat value Te. In this study, T0 is reduced to the critical value Tc0 for a synergetic second-order transition (see section 3). When the modulus defect 33401-6 Self-organization of adatom adsorption structure at interaction with the tip of dynamic force microscope is taken into account, the temperature T0 = 1+ Te −1 1+ gε2 0 , (33) whose value is defined by a minimum position of the dependence (30), is realized. In accordance with equations (32) and (33), the quantity T0 monotonously decays from the value Tm = 1+ T 0 c −1 1+ g ( εc 0 )2 , εc 0 = ( 2gθ )−1 [ g ( T 0 c −1−α2 ) −θ ] (34) at Te = T 0 c , to 1 at Te →∞. As shown in figure 3, the stationary temperature T0 = Te in the range from 0 to Tc0. The jump down occurs at Te = Tc0, following which the value T0 smoothly decreases. If the quantity Te then decays, the steady-state temperature T0 increases. At the point T 0 c [equation (31)], the T0 vs Te dependence has a jump from Tm [equation (34)] up to T 0 c . For Te < T 0 c , the steady-state temperature T0 is also equal to Te. Figure 3. Dependence of the steady-state value of the sample temperature T0 on the temperature Te at parameters of figure 1. Since T 0 c > 1, the maximal sample temperature (34) is lower than the minimal temperature of the AFM tip (31). As shown in figure 3, at Te > T 0 c , the stationary temperature T0 of the sample is smaller than Te. 5. Summary In accordance with the analysis presented above, the formation of an adatom adsorption structure is caused by self-organization of shear components of the strain and by the stress fields, on the one hand, and by the sample temperature, on the other hand. Here, the strain ε acts as the order parameter, the conjugate field is reduced to the stress σ, and the temperature T is the control parameter. The cause for self-organization is the positive feedback of T and ε on σ [see equation (13)]. According to equa- tions (2) and (4), it is caused by the temperature dependence of the shear modulus. With an allowance for the effective shear modulus vs strain dependence, we obtain expressions for temperatures corre- sponding to the absolute instability of the adatom configuration T 0 c [equation (31)] and its stability limit Tc0 [equation (23)]. A real thermodynamic transition temperature can be determined from the equality V (0) =V (ε0) of potentials in different phases and it is in the (T 0 c ,Tc0) region. According to equation (23), systems predisposed to the formation of adatom structure have large typical G0 and small effective Gε values of shear modulus. 33401-7 A.V. Khomenko The present study is principally different from [30]. In particular, the basic parameters and equations are different. For example, the order parameter is the strain, the shear modulus depends on temperature (4), the derivation of equation for temperature (10) differs and so on. Therefore, the resultant equations and figures are different from the ones in [30]. Thus, solid-liquid transition of an ultrathin lubricant film and self-organization of adatoms on the semiconductor surface in contact with the tip of the dynamic force microscope are described only by a similar approach but with many different aspects. At a choice of real parameters, there is a difficulty connected primarily with the following features which are common with the ultrathin lubricant film [30, 45, 48, 49]. The properties of adatom layers due to their small thickness do not coincide with the properties of volume materials. They are characterized by various values of elastic constants, density, heat conductivity, etc. The adatom structure temperature is also an effective quantity, and it can essentially fluctuate, since the adatoms number is limited, and these fluctuations lead to transitions between states [31, 50]. Therefore, we are restricted by a description of the qualitative system behavior and all parameters are transformed into dimensionless form. Acknowledgements The basis of this study method was founded in the joint papers with my teacher Prof. A.I. Olemskoi cited here. I thank Dr. Bo N.J. Persson for the invitation, hospitality, helpful comments and suggestions during my stay in the Forschungszentrum Jülich (Germany). I am grateful to him and to the organizers of the conference “Joint ICTP-FANAS Conference on Trends in Nanotribology” (12–16 September 2011, Miramare, Trieste, Italy) for the invitation and financial support for participation, duringwhich this work was initiated. The work was supported by the grant of the Ministry of Education and Science of Ukraine “Modelling of friction of metal nanoparticles and boundary liquid films which interact with atomically flat surfaces” (No. 0112U001380) and by the grant for a research visit to the Forschungszentrum Jülich (Germany). I am thankful to Dr. Boris Lorenz for an attentive reading and correction of this article. References 1. Fundamentals of Friction and Wear on the Nanoscale, Gnecco E., Meyer E. (Eds.), Springer, Berlin, 1 edn., 2007. 2. Garcia R., Amplitude Modulation Atomic Force Microscopy, Wiley-VCH Verlag GmbH and Co. KGaA, Weinheim, Germany, 2010. 3. Garcia R., Pérez R., Surf. Sci. Rep., 2002, 47, No. 6–8, 197; doi:10.1016/S0167-5729(02)00077-8. 4. Szlufarska I., Chandross M., Carpick R.W., J. Phys. D: Appl Phys, 2008, 41, No. 12, 123001; doi:10.1088/0022-3727/41/12/123001. 5. Hofer W., Foster A., Shluger A., Rev. Mod. Phys., 2003, 75, No. 4, 1287; doi:10.1103/RevModPhys.75.1287. 6. Giessibl F., Rev. Mod. Phys., 2003, 75, No. 3, 949; doi:10.1103/RevModPhys.75.949. 7. Pogrebnjak A.D., Shpak A.P., Azarenkov N.A., Beresnev V.M., Phys.-Usp., 2009, 52, No. 1, 29; doi:10.3367/UFNe.0179.200901b.0035. 8. Wang X., Kunc K., Loa I., Schwarz U., Syassen K., Phys. Rev. B, 2006, 74, 134305; doi:10.1103/PhysRevB.74.134305. 9. Guo W., Zhu C.Z., Yu T.X., Woo C.H., Zhang B., Dai Y.T., Phys. Rev. Lett., 2004, 93, 245502; doi:10.1103/PhysRevLett.93.245502. 10. Yong C., Kendall K., Smith W., Philos. Trans. R. Soc. London, Ser. A, 2004, 362, No. 1822, 1915; doi:10.1098/rsta.2004.1423. 11. Oyabu N., Pou P., Sugimoto Y., Jelinek P., Abe M., Morita S., Pérez R., Custance O., Phys. Rev. Lett., 2006, 96, 106101; doi:10.1103/PhysRevLett.96.106101. 12. Khomenko A.V., Prodanov N.V., Carbon, 2010, 48, No. 4, 1234; doi:10.1016/j.carbon.2009.11.046. 13. Prodanov N.V., Khomenko A.V., Surf. Sci., 2010, 604, No. 7–8, 730; doi:10.1016/j.susc.2010.01.024. 14. Khomenko A.V., Prodanov N.V., J. Phys. Chem. C, 2010, 114, 19958; doi:10.1021/jp108981e. 15. Pogrebnyak A.D., Ponomarev A.G., Shpak A.P., Kunitskii Y.A., Phys.-Usp., 2012, 55, No. 3, 270; doi:10.3367/UFNe.0182.201203d.0287. 16. Langewisch G., Kamiński W., Braun D.A., Moller R., Fuchs H., Schirmeisen A., Pérez R., Small, 2012, 8, No. 4, 602; doi:10.1002/smll.201101919. 17. Campbellová A., Ondrácek M., Pou P., Pérez R., Klapetek P., Jelinek P., Nanotechnology, 2011, 22, No. 29, 295710; doi:10.1088/0957-4484/22/29/295710. 33401-8 http://dx.doi.org/10.1016/S0167-5729(02)00077-8 http://dx.doi.org/10.1088/0022-3727/41/12/123001 http://dx.doi.org/10.1103/RevModPhys.75.1287 http://dx.doi.org/10.1103/RevModPhys.75.949 http://dx.doi.org/10.3367/UFNe.0179.200901b.0035 http://dx.doi.org/10.1103/PhysRevB.74.134305 http://dx.doi.org/10.1103/PhysRevLett.93.245502 http://dx.doi.org/10.1098/rsta.2004.1423 http://dx.doi.org/10.1103/PhysRevLett.96.106101 http://dx.doi.org/10.1016/j.carbon.2009.11.046 http://dx.doi.org/10.1016/j.susc.2010.01.024 http://dx.doi.org/10.1021/jp108981e http://dx.doi.org/10.3367/UFNe.0182.201203d.0287 http://dx.doi.org/10.1002/smll.201101919 http://dx.doi.org/10.1088/0957-4484/22/29/295710 Self-organization of adatom adsorption structure at interaction with the tip of dynamic force microscope 18. Barth C., Foster A.S., Henry C.R., Shluger A.L., Adv. Mater., 2011, 23, No. 4, 477; doi:10.1002/adma.201002270. 19. Proksch R., Kalinin S.V., Nanotechnology, 2010, 21, No. 45, 455705; doi:10.1088/0957-4484/21/45/455705. 20. Dieška P., Štich I., Phys. Rev. B, 2009, 79, 125431; doi:10.1103/PhysRevB.79.125431. 21. Pokropivnyi A.V., Tech. Phys. Lett., 2000, 26, No. 11, 967; doi:10.1134/1.1329686. 22. Dieška P., Štich I., Pérez R., Phys. Rev. Lett., 2005, 95, 126103; doi:10.1103/PhysRevLett.95.126103. 23. Sahagún E., Sáenz J.J., Phys. Rev. B, 2012, 85, 235412; doi:10.1103/PhysRevB.85.235412. 24. Lange M., van Vorden D., Moller R., Beilstein J. Nanotech., 2012, 3, 207; doi:10.3762/bjnano.3.23. 25. Barsoum M.W., Murugaiah A., Kalidindi S.R., Zhen T., Phys. Rev. Lett., 2004, 92, 255508; doi:10.1103/PhysRevLett.92.255508. 26. Haken H., Synergetics. An introduction. Nonequilibrium phase transitions and self-organization in physics, chemistry, and biology, Springer, Berlin, 3 edn., 1983. 27. Landau L.D., Lifshitz E.M., Course of Theoretical Physics, Vol.5: Statistical Physics, Butterworth, London, 1999. 28. Lifshits E.M., Pitaevskii L.P., Course of Theoretical Physics, Vol.10: Physical Kinetics, Pergamon Press, Oxford, 1 edn., 1981. 29. Olemskoi A.I., Khomenko A.V., J. Exp. Theor. Phys., 1996, 83, No. 6, 1180. 30. Khomenko A.V., Yushchenko O.V., Phys. Rev. E, 2003, 68, 036110; doi:10.1103/PhysRevE.68.036110. 31. Khomenko A.V., Phys. Lett. A, 2004, 329, No. 1–2, 140; doi:10.1016/j.physleta.2004.06.091. 32. Rheology, Eirich F. (Ed.), Academic Press, New York, 1960. 33. Landau L.D., Lifshitz E.M., Course of Theoretical Physics, Vol.7: Theory of Elasticity, Pergamon Press, New York, 3 edn., 1986. 34. Mate C., IBM J. Res. Dev., 1995, 39, No. 6, 617; doi:10.1147/rd.396.0617. 35. Khomenko A.V., Prodanov N.V., Condens. Matter Phys., 2008, 11, No. 4, 615; doi:10.5488/CMP.11.4.615. 36. Gómez C.J., Garcia R., Ultramicroscopy, 2010, 110, No. 6, 626; doi:10.1016/j.ultramic.2010.02.023. 37. Landau L.D., Khalatnikov I.M., Dokl. Akad. Nauk SSSR, 1954, 96, 469, (see also: Collected Papers of L.D. Landau, edited by D. ter Haar. Pergamon, London (1965)). 38. Berthier L., Biroli G., Rev. Mod. Phys., 2011, 83, No. 2, 587; doi:10.1103/RevModPhys.83.587. 39. Havranek A., Marvan M., Ferroelectrics, 1996, 176, 25; doi:10.1080/00150199608223597. 40. Olemskoi A.I., Khomenko A.V., Tech. Phys., 2000, 45, 672; doi:10.1134/1.1259700. 41. Olemskoi A.I., Khomenko A.V., Tech. Phys., 2000, 45, 677; doi:10.1134/1.1259702. 42. Khomenko A.V., Lyashenko I.A., Condens. Matter Phys., 2006, 9, No. 4(48), 695; doi:10.5488/CMP.9.4.695. 43. Olemskoi A.I., Khomenko A.V., Kharchenko D.O., Physica A, 2003, 323, 263; doi:10.1016/S0378-4371(02)01991-X. 44. Olemskoi A.I., Khomenko A.V., Phys. Rev. E, 2001, 63, 036116; doi:10.1103/PhysRevE.63.036116. 45. Khomenko A.V., Lyashenko Y.A., Tech. Phys., 2010, 55, No. 1, 26; doi:10.1134/S1063784210010056. 46. Metlov L.S., Phys. Rev. E, 2010, 81, 051121; doi:10.1103/PhysRevE.81.051121. 47. Metlov L.S., Phys. Rev. Lett., 2011, 106, 165506; doi:10.1103/PhysRevLett.106.165506. 48. Khomenko A.V., Lyashenko I.A., J. Phys. Stud., 2007, 11, No. 3, 268. 49. Khomenko A.V., Lyashenko I.A., Phys. Lett. A, 2007, 366, No. 1–2, 165; doi:10.1016/j.physleta.2007.02.010. 50. Khomenko A.V., Lyashenko I.A., Tech. Phys., 2005, 50, No. 11, 1408; doi:10.1134/1.2131946. 33401-9 http://dx.doi.org/10.1002/adma.201002270 http://dx.doi.org/10.1088/0957-4484/21/45/455705 http://dx.doi.org/10.1103/PhysRevB.79.125431 http://dx.doi.org/10.1134/1.1329686 http://dx.doi.org/10.1103/PhysRevLett.95.126103 http://dx.doi.org/10.1103/PhysRevB.85.235412 http://dx.doi.org/10.3762/bjnano.3.23 http://dx.doi.org/10.1103/PhysRevLett.92.255508 http://dx.doi.org/10.1103/PhysRevE.68.036110 http://dx.doi.org/10.1016/j.physleta.2004.06.091 http://dx.doi.org/10.1147/rd.396.0617 http://dx.doi.org/10.5488/CMP.11.4.615 http://dx.doi.org/10.1016/j.ultramic.2010.02.023 http://dx.doi.org/10.1103/RevModPhys.83.587 http://dx.doi.org/10.1080/00150199608223597 http://dx.doi.org/10.1134/1.1259700 http://dx.doi.org/10.1134/1.1259702 http://dx.doi.org/10.5488/CMP.9.4.695 http://dx.doi.org/10.1016/S0378-4371(02)01991-X http://dx.doi.org/10.1103/PhysRevE.63.036116 http://dx.doi.org/10.1134/S1063784210010056 http://dx.doi.org/10.1103/PhysRevE.81.051121 http://dx.doi.org/10.1103/PhysRevLett.106.165506 http://dx.doi.org/10.1016/j.physleta.2007.02.010 http://dx.doi.org/10.1134/1.2131946 A.V. Khomenko Самоорганiзацiя структури адсорбованих адатомiв при взаємодiї iз зондом динамiчного силового мiкроскопа О.В. Хоменко1,2 1 Кафедра моделювання складних систем, Сумський державний унiверситет, вул. Римського-Корсакова, 2, 40007 Суми, Україна 2 Iнститут Петера Грюнберга-1, Дослiдницький центр Юлiха, 52425 Юлiх, Нiмеччина Формування структури адсорбованих адатомiв при дослiдженнi в режимi динамiчної силової мiкроскопiї представлено як результат спонтанної появи зсувної деформацiї в результатi зовнiшнього надкритич- ного нагрiвання. Цей перехiд описується рiвнянням Кельвiна-Фойгта для в’язкопружного середовища, релаксацiйним рiвнянням Ландау-Халатнiкова для зсувних напружень та релаксацiйним рiвнянням для температури. Показано, що цi рiвняння формально збiгаються iз синергетичною системою Лоренца, де зсувна деформацiя вiдiграє роль параметра порядку, спряжене поле зводиться до напружень, та темпера- тура є керувальним параметром. В рамках адiабатичного наближення знайденi стацiонарнi значення цих величин. Враховуючи залежнiсть модуля зсуву зразка вiд деформацiї, формування конфiгурацiї адсорбо- ваних адатомiв описано як перехiд першого роду. Критична температура зонда лiнiйно зростає з ростом ефективного значення модуля зсуву зразка i зменшується при зростаннi його характерного значення. Ключовi слова: фазовий перехiд, реологiя, пластичнiсть, деформацiя, напруження, атомно-силова мiкроскопiя 33401-10 Introduction Basic equations Continuous transition Effect of deformational defect of modulus Summary