1: %\documentstyle[preprint,aps]{revtex}
2: \documentstyle[eqsecnum,aps]{revtex}
3: \def\btt#1{{\tt$\backslash$#1}}
4: \begin{document}
5: \draft \preprint{HEP/123-qed}
6: \title{Path Integral of the Holstein Model with a $\phi^4$ on site
7: potential}
8: \author{ Marco Zoli }
9: \address{Istituto Nazionale Fisica della Materia - Dipartimento di Fisica
10: \\ Universit\'a di Camerino, 62032, Italy. - marco.zoli@unicam.it}
11:
12: \date{\today}
13:
14: \maketitle
15: \begin{abstract}
16: We derive the path integral of the semiclassical, one dimensional
17: anharmonic Holstein model assuming that the electron motion takes
18: place in a bath of non linear oscillators with quartic on site
19: hard (and soft) potentials. The interplay between {\it e-ph}
20: coupling and anharmonic force constant is analysed both in the
21: adiabatic and antiadiabatic regime. In the latter we find much
22: larger anharmonic features on the thermodynamic properties of low
23: energy oscillators. Soft on site potentials generate attractive
24: centres at large amplitude oscillator paths and contribute to the
25: anomalous shape of the {\it heat capacity over temperature} ratio
26: in the intermediate to low $T$ range. This anharmonic lattice
27: effect is superimposed to the purely electronic contribution
28: associated to a temperature dependent hopping with variable range
29: inducing local disorder in the system.
30: \end{abstract}
31:
32: \pacs{PACS: 31.15.Kb, 71.38.-k, 63.20.Ry }
33:
34: \narrowtext
35: \section*{I. Introduction}
36:
37: There is at present a large interest on the effects of a strong
38: electron-phonon coupling in a number of systems ranging from dimer
39: molecular junctions \cite{kaat} to carbon nanotubes \cite{mah},
40: from organic molecular crystals \cite{han}, to DNA \cite{alexan}
41: and cuprate superconductors \cite{sch,lan}. Several theoretical
42: studies have focussed on the interplay between e-ph coupling and
43: non linearities in the framework of the Holstein model
44: \cite{chris,voulga} investigating the phase diagram both in the
45: adiabatic \cite{zieg} and the antiadiabatic regime. Mainly in the
46: latter anharmonic effects are believed to be large \cite{pietro}
47: thus offering a picture to explain the high $T_c$ of binary alloys
48: such as superconducting $MgB_2$ with a small Fermi energy and
49: sizeable {\it e-ph} coupling \cite{cava}.
50:
51: The path integral formalism provides a powerful method to study
52: quantum systems in which a particle is non linearly coupled to the
53: environment \cite{fehi,kleinert,eck}. A previous path integral
54: analysis \cite{io5} has pointed out how the phonon dispersion,
55: that has to be taken into account in the computation of the ground
56: state properties of the Holstein Hamiltonian
57: \cite{holst1,raedt2,io98}, induces non local {\it e-ph}
58: correlations which renormalize downwards the effective coupling
59: and ultimately broaden the size of the polaronic quasiparticle.
60: This explains why the polaron mass in a dispersive Holstein model
61: \cite{atin} is lighter than in a dispersionless model
62: \cite{alekor}. Also the thermodynamics of the Holstein Hamiltonian
63: can be computed within a dispersive model which accounts for the
64: lattice structure \cite{io5}.
65:
66: The Holstein diatomic molecular model was originally cast
67: \cite{holst} in the form of a discrete non linear Schr\"odinger
68: equation for electrons whose probability amplitude at a molecular
69: site depends on the interatomic vibration coordinates. The non
70: linearities are tuned by the {\it e-ph} coupling \cite{kenkre}
71: whose strength drives the crossover between a large and a small
72: polaron state according to the degree of adiabaticity and the
73: dimensionality of the system \cite{trug}.
74:
75: In the Holstein model, the phonon thermodynamics is not affected
76: by {\it e-ph} induced anharmonicities \cite{io4}. This follows
77: from the fact that the Holstein perturbing source current is local
78: in time and it does not depend on the electron path coordinate. As
79: a consequence, in the total partition function, electron and
80: lattice degrees of freedom are disentangled and the latter can be
81: integrated out analytically as far as a harmonic lattice model is
82: assumed. However non linearities may arise in Holstein like
83: systems also by virtue of on site potentials dependent on the
84: lattice structure and, in principle, independent of the {\it e-ph}
85: coupling. We are thus led to investigate the thermodynamics of the
86: anharmonic Holstein model with a quartic on site potential which
87: may be repulsive or partly attractive according to the sign of the
88: force constant and the amplitude of the lattice displacement
89: paths. The path integral approach permits to monitor the physical
90: properties for any value of the coupling strengths \cite{feynman}.
91: The presence of a $\phi^4$ potential may in turn affect also the
92: {\it e-ph} interactions and sinergically interfere on the
93: equilibrium properties of the system. This is the focus of the
94: present paper. Section II presents the Hamiltonian model while the
95: path integral method is briefly described in Section III. The
96: derivation of the total partition function of the system is
97: presented in Section IV and Section V contains the discussion of
98: the physical results. The conclusions are drawn in Section VI.
99:
100: \section*{II. The Anharmonic Holstein Model}
101:
102: We consider the one dimensional anharmonic Holstein Hamiltonian
103: consisting of: i) one electron hopping term, ii) an interaction
104: which couples the electronic density ($f_{\bf l}^{\dag}f_{\bf l}$)
105: to the lattice displacement $u_l$ at the {l}-site, iii) a bath of
106: $N$ identical dispersionless anharmonic oscillators with mass $M$
107: and frequency $\omega$ :
108:
109: \begin{eqnarray}
110: & &H =\, H^e + H^{e-ph} + H^{ph} \nonumber
111: \\
112: & &H^e =\, - t \sum_{<{l, m}>} f_{l}^{\dag} f_{m} \nonumber
113: \\
114: & &H^{e-ph}=\, {\bar g} \sum_{l=1}^Nu_l f^{\dag}_l f_{l} \,
115: \nonumber
116: \\
117: & &H^{ph}=\, {M \over 2} \sum_{l=1}^N \bigl(\dot{u}_l^2 + \omega^2
118: u_l^2 \bigr) + {{\delta} \over 4} \sum_{l=1}^N u_l^4 \, \nonumber
119: \\ & &{\bar g}=\,g \sqrt{2M \omega}
120: \end{eqnarray}
121:
122: the sum $<,>$ is over $z$ nearest neighbors and $t$ is the tight
123: binding overlap integral. $g$ is the {\it e-ph} coupling in units
124: of $\hbar \omega$. Choosing the atomic mass $M$ of order $10^4$
125: times the electron mass, we get ${\bar g} \simeq \, 1.1456 \times
126: g \sqrt{2 \hbar \omega}{}\, \bigl[ meV \AA^{-1} \bigr]$ where
127: $\hbar \omega$ is given in $meV$. $\delta$, in units $meV
128: \AA^{-4}$, controls the strength of the non linearities and
129: determines whether the on site potential $V(u_l)=\, {M}\omega^2
130: u_l^2/2 + {{\delta}}u_l^4/4$ is hard ($\delta > 0$) or soft
131: ($\delta < 0$). In the latter case, $V(u_l)$ attains the maximum
132: at $u_l^2=\,M\omega^2/|\delta|$ and the inflection point occurs at
133: $u_l^2=\,M\omega^2/3|\delta|$. The condition $|\delta|u_l^2 > 2
134: M\omega^2$ yields an attractive on site potential. Then, the range
135: of the atomic path amplitudes generating attractive scattering
136: centres depends on the value of the anharmonic force constant. For
137: $|\delta| > 2 M\omega^2$ the potential becomes attractive for a
138: portion of large amplitude atomic paths while small amplitude
139: paths weigh the repulsive range.
140:
141: In the following computation of the electron path integral coupled
142: to the anharmonic oscillator, after setting the potential
143: parameters, we select at any temperature the class of atomic paths
144: which mainly contribute to the euclidean action. As the
145: distribution of the path amplitudes has a cutoff on the scale of
146: the lattice constant, say i.e. $u_l^2 < 1 \AA^2$, the on site
147: potentials are always bound from below also in the attractive
148: cases.
149:
150: \section*{III. The Path Integral Method}
151:
152: The Holstein Hamiltonian in (1) can be mapped onto the time scale
153: according to space-time mapping techniques extensively described
154: in previous works \cite{io5,hamann,io3} and hereafter outlined.
155:
156: Defining ${x}(\tau)$ and ${y}(\tau')$ as the electron coordinates
157: at the ${l}$ and ${m}$ lattice sites respectively, $H^e$ in (1)
158: transforms into
159:
160: \begin{equation}
161: H^e(\tau,\tau')=\, -{t} \bigl( f^{\dag}({x}(\tau)) f({y}(\tau')) +
162: f^{\dag}({y}(\tau')) f({x}(\tau)) \bigr)
163: \end{equation}
164:
165: where $\tau$ and $\tau'$ are continuous variables $\bigl( \in [0,
166: \beta]\bigr)$, with $\beta$ being the inverse temperature. After
167: setting $\tau'=\,0$, ${y}(0) \equiv 0$ and taking the thermal
168: averages for the electron operators over the ground state of the
169: Hamiltonian one gets the average electron hopping energy per
170: lattice site:
171:
172: \begin{eqnarray}
173: & &h^e(x(\tau)) \equiv {{<H^e(\tau)>} \over N}= \, - {t}\Bigl(G[-{
174: x}(\tau), -\tau ] + G[{x}(\tau), \tau ]\Bigr)\, \nonumber \\
175: \end{eqnarray}
176:
177: where, $G[{x}(\tau), \tau ]$ is the electron propagator at finite
178: temperature.
179:
180: By treating the lattice displacements in (1) as $\tau -$ dependent
181: classical variables, $u_l \rightarrow u(\tau)$, we obtain from
182: $H^{e-ph}$ in (1) the averaged {\it e-ph} energy per lattice site
183: which is identified as the perturbing source current $j(\tau)$ in
184: the path integral method:
185:
186: \begin{eqnarray}
187: j(\tau) \equiv {{<H^{e-ph}(\tau)>} \over N}= \,{\bar g} u(\tau)
188: \end{eqnarray}
189:
190: As the Hamiltonian model assumes a set of identical oscillators we
191: study the path integral for the electron coupled to a single
192: anharmonic oscillator of the bath. The path integral reads:
193:
194: \begin{eqnarray}
195: & &<{x}(\beta)|{x}(0)> =\, \int D{x}(\tau) exp\Biggl[-
196: \int_0^{\beta}d\tau E(x(\tau))\Biggr] \nonumber
197: \\ &\times& \int Du(\tau) exp\Biggl[- \int_0^{\beta}
198: d\tau O(u(\tau)) \Biggr] \,\nonumber
199: \\
200: & &E(x(\tau))=\,{{m_e} \over 2} \dot{{x}}^2(\tau) + h^e(x(\tau))
201: \,\nonumber
202: \\
203: & &O(u(\tau))=\, {M \over 2} \bigl( \dot{u}^2(\tau) + \omega^2
204: u^2(\tau) \bigr) + {{\delta} \over 4} u^4(\tau) +
205: j(\tau)\,\nonumber
206: \\
207: \end{eqnarray}
208:
209: where the kinetic term ($m_e$ is the electron mass) is normalized
210: by the functional measure of integration over the electron paths.
211:
212: Since the electron hopping does not induce a shift of the
213: oscillator coordinate the Holstein {\it e-ph} interactions are
214: local in time and, in the semiclassical treatment, the source
215: current $j(\tau)$ is independent of the electron path. As a
216: consequence oscillator and electron coordinates appear
217: disentangled in (5) while the coupling occurs through the
218: parameter $\bar{g}$.
219:
220: \section*{IV. The Partition Function}
221:
222: The quantum statistical partition function $Z_T$ is derived by
223: integrating (5) after imposing periodicity conditions, $\beta$ is
224: the period, both on the electron and oscillator paths:
225:
226: \begin{eqnarray}
227: Z_T &=&\,\int dx <{x}(\beta)|{x}(0)> =\, Z_{el} \times Z_{osc}
228: \nonumber
229: \\
230: Z_{el} &=&\, \oint Dx(\tau) exp \Bigl[- \int_0^{\beta}d\tau
231: E(x(\tau)) \Bigr] \nonumber
232: \\
233: Z_{osc} &=&\, \oint Du(\tau) exp\Bigl[- \int_0^{\beta} d\tau
234: O(u(\tau)) \Bigr]
235: \,\nonumber \\
236: \end{eqnarray}
237:
238: where $\oint Dx(\tau)$ and $\oint Du(\tau)$ are the functional
239: measures of integration.
240:
241: The electronic contribution $Z_{el}$ is computed by expanding the
242: paths in Fourier components
243:
244: \begin{eqnarray}
245: x(\tau)&=&\,x_o + \sum_{m=1}^{M_F}\bigl( r_m \cos(\nu_m \tau) +
246: s_m \sin(\nu_m \tau) \bigr)\, \nonumber
247: \\
248: &&r_m =\,2\Re x_m \, \nonumber
249: \\
250: &&s_m =\, -2\Im x_m \, \nonumber
251: \\
252: &&\nu_m =\,2m\pi/\beta \, \nonumber
253: \\
254: \end{eqnarray}
255:
256:
257: and taking the following measure of integration
258:
259: \begin{eqnarray}
260: && \oint Dx(\tau) \equiv {{\sqrt{2}} \over {(2
261: \lambda_{m_e})^{(2M_F+1)}}} \int_{-\infty}^{\infty}{dx_o} \,
262: \nonumber
263: \\
264: &\times&
265: \prod_{m=1}^{M_F} (2\pi m)^2 \int_{-\infty}^{\infty} dr_m
266: \int_{-\infty}^{\infty} ds_m \, \nonumber
267: \\
268: & &\lambda_{m_e}=\,\sqrt{\pi \hbar^2 \beta/m_e}\, \nonumber
269: \\
270: \end{eqnarray}
271:
272: The cutoffs over the Fourier coefficient integrations have to
273: ensure proper normalization of the kinetic term in absence of
274: hopping processes. Thus $Z_{el}$ transforms into
275:
276: \begin{eqnarray}
277: Z_{el}&\simeq& \,{{\sqrt{2}} \over {(2 \lambda_{m_e})^{(2M_F+1)}}}
278: \int_{-\Lambda/2}^{\Lambda/2}{d{x}_o} \prod_{m=1}^{M_F}
279: \int_{-\Lambda}^{\Lambda} d{r}_m \int_{-\Lambda}^{\Lambda} d{s}_m
280: \, \nonumber
281: \\ &\times& \exp\biggl(- {{\pi^{3}} \over \lambda^2_{m_e}} \sum_{m=1}^{M_F}
282: m^2({r}^2_m + {s}^2_m) - \int_0^{\beta}d\tau h^e(x(\tau)) \biggr)
283: \, \nonumber
284: \\
285: \end{eqnarray}
286:
287: with $\Lambda \propto \lambda_{m_e}$ \cite{io3} indicating that
288: large amplitude electron paths have to be selected at low
289: temperatures where the quantum effects are larger. Two Fourier
290: components, $M_F=\,2$, suffice to attain stable results as
291: $h^e(x(\tau))$ depends smoothly on the electron path. The hopping
292: term accounts for the deviation from the Gaussian behavior.
293: Numerical analysis shows that, for any choice of path parameters,
294: $h^e(x(\tau))$ decreases by decreasing the temperature but its
295: overall contribution to the electron action is substantial also at
296: low $T$.
297:
298: Let's focus now on the anharmonic oscillator term $Z_{osc}$. The
299: oscillator path is expanded in $N_F$ Fourier components
300:
301: \begin{eqnarray}
302: u(\tau)=\,u_o + \sum_{n=1}^{N_F}\bigl( a_n \cos(\omega_n \tau) +
303: b_n \sin(\omega_n \tau) \bigr)\, \nonumber
304: \\
305: \end{eqnarray}
306:
307: with Matsubara frequencies $\omega_n=\,2n\pi/\beta$ and
308: coefficients $a_n\equiv \Re u_n$, $b_n\equiv -\Im u_n$ satisfying
309: the conditions $a_n =\,a_{-n}$ and $b_n =\,-b_{-n}$. The latter
310: are consistent with the choice of real paths and simplify the
311: following $\tau$ integration of the on site potential.
312:
313: Note that the periodicity property, $u(\tau) =\,u(\tau + \beta)$,
314: would be fulfilled also taking the very $a_n$ coefficients in (10)
315: \cite{fey}. However such a choice would not permit to fit with
316: accuracy the harmonic oscillator partition function which is known
317: exactly: $Z_h =\,\bigl[2 \sinh(\beta\omega/2)\bigr]^{-1}$. Infact,
318: in the path integral method, the harmonic partition function
319: $Z^{PI}_h$ reads
320:
321: \begin{eqnarray}
322: Z^{PI}_h =\, \oint Du(\tau) exp\Biggl[- \int_0^{\beta} d\tau
323: \Bigl[ {M \over 2} \bigl( \dot{u}^2(\tau) + \omega^2 u^2(\tau)
324: \bigr) \Bigr] \Biggr]\, \nonumber
325: \\
326: \end{eqnarray}
327:
328: and taking the functional measure
329:
330: \begin{eqnarray}
331: \oint Du(\tau) \equiv & & {{\sqrt{2}} \over {(2
332: \lambda_M)^{(2N_F+1)}}} \int_{-\infty}^{\infty}{du_o} \,
333: \nonumber
334: \\
335: &\times& \prod_{n=1}^{N_F} (2\pi n)^2 \int_{-\infty}^{\infty}
336: da_n \int_{-\infty}^{\infty} db_n \, \nonumber
337: \\
338: \end{eqnarray}
339:
340: with $\lambda_M=\,\sqrt{\pi \hbar^2 \beta/M}$, one gets from
341: (10)-(12):
342:
343: \begin{eqnarray}
344: Z^{PI}_h=\, {1 \over {\beta \omega}} \prod_{n=1}^{N_F}
345: {{(2n\pi)^2} \over {{(2n\pi)^2 + (\beta \omega)^2}}} \, \nonumber
346: \\
347: \end{eqnarray}
348:
349: Instead, dropping the $b_n$ terms in (10) and (12), one would get
350: the square root of the product series in (13) which does not yield
351: a reliable fit of $Z_h$ even for large $N_F$. Note that at high
352: $T$, the condition $2n\pi \gg \beta \omega$ is fulfilled for small
353: integers $n$, hence the main contribution to $Z^{PI}_h$ is given
354: by the ${1 /{\beta \omega}}$ factor which stems from the
355: $\int{du_o}$ in (12). This is consistent with the expectation that
356: high $T$ paths are well approximated by their $\beta$-averaged
357: value $u_o$ whereas fluctuation effects become increasingly
358: relevant towards the low $T$ regime in which $N_F$ rapidly grows.
359: $N_F$ clearly varies also with the oscillator energy while the
360: shape of $N_F(\omega,T)$ may differ according to the harmonic
361: function ($Z_h$, harmonic free energy or specific heat) one
362: chooses to fit.
363:
364: The anharmonic partition function $Z_{osc}$ in (6) can be worked
365: out analytically using (10),(12) and performing the time
366: integration of the oscillator functional $O(u(\tau))$. This
367: permits to get an insight into the role of the non linear terms.
368: The lenghty calculation yields:
369:
370:
371: \begin{eqnarray}
372: &&Z_{osc}=\, {{\sqrt{2}} \over {(2 \lambda_M)^{(2N_F+1)}}}
373: \int_{-\infty}^{\infty}{du_o} \exp\bigl( -\beta \bar g u_o -
374: \kappa u_o^2 -\beta\delta u_o^4/4 \bigr)\, \nonumber \\ && \times
375: \prod_{n=1}^{N_F} (2\pi n)^2 \int_{-\infty}^{\infty} da_n
376: \exp\Biggl[-\sum_{n=1}^{N_F} \Bigl[ (\gamma_n + 3\beta\delta
377: u_o^2/4)a_n^2 \nonumber
378: \\
379: &&+ {{\beta\delta u_o} \over 2} \sum_{m=1}^{N_F}c({n,m}) +
380: {{\beta\delta} \over {16}} \sum_{m,p=1}^{N_F}d({n,m,p}) \Bigr]
381: \Biggr]\, \nonumber \\ && \times \int_{-\infty}^{\infty} db_n
382: \exp\Biggl[-\sum_{n=1}^{N_F} \Bigl[ (\gamma_n + 3\beta\delta
383: u_o^2/4)b_n^2 \nonumber
384: \\
385: &&+ {{3\beta\delta u_o} \over 2} \sum_{m=1}^{N_F}e({n,m}) +
386: {{\beta\delta} \over {16}} \sum_{m,p=1}^{N_F}\bigl({{6}}f({n,m,p})
387: - g({n,m,p}) \bigr) \Bigr] \Biggr]\, \nonumber \\ \nonumber
388: \\
389: &&\kappa=\,\pi
390: (\beta\omega)^2/2\lambda_M^2 \, \nonumber
391: \\
392: && \gamma_n=\, \pi \bigl((2\pi n)^2 + (\beta\omega)^2 \bigr)/4
393: \lambda_M^2 \, \nonumber
394: \\
395: && c({n,m})=\,a_n a_m (a_{n+m}+a_{n-m}) \nonumber
396: \\
397: && d({n,m,p})=\,a_n a_m a_p
398: (a_{n+m+p}+a_{n-m+p}+a_{p-n-m}+a_{p-n+m}) \nonumber
399: \\
400: && e({n,m})=\,a_n b_m (b_{n+m}- b_{n-m}) \nonumber
401: \\
402: && f({n,m,p})=\,a_n a_m b_p
403: (b_{n+m+p}+b_{n-m+p}+b_{p-n-m}+b_{p-n+m}) \nonumber
404: \\
405: && g({n,m,p})=\,b_n b_m b_p
406: (b_{n+m+p}-b_{n-m+p}+b_{p-n-m}-b_{p-n+m}) \nonumber
407: \\
408: \end{eqnarray}
409:
410: In $c({n,m})$ the coefficients $a_{j}$ $(j=\, n+m, n-m)$ vanish
411: if: $j \leq 0$ or $j > N_F$. In $d({n,m,p})$, $a_k \neq 0 \,\,
412: (k=\,n+m+p, \,\, n-m+p, \,\, p-n-m, \,\, p-n+m)
413: \Longleftrightarrow 1 \leq k \leq N_F$. Analogous conditions hold
414: for the coefficients $b_{j\, (k)}$ in the $e({n,m})$, $f({n,m,p})$
415: and $g({n,m,p})$ functions.
416:
417: Note that the effective {\it e-ph} coupling $\bar{g}$ is
418: associated only to the $\tau$-independent component $u_o$, that is
419: to the $\beta$-averaged displacement path \cite{io5}. This follows
420: from the fact that the perturbing source current is non retarded
421: in the Holstein model as a consequence of the local nature of the
422: e-ph interaction.
423:
424: The quartic potential induces a strong mixing of the Fourier
425: components of the path that highly complicates the numerical
426: problem. Thus the value of $N_F$ appears to be crucial in the
427: computation. We determine $N_F(\omega,T)$ by fitting (with an
428: accuracy of $2 \cdot 10^{-2}$) the exact harmonic free energy
429: ($F_h=\,-\ln(Z_h)/\beta$), through the path integral harmonic free
430: energy $F_h^{PI}$ obtained from (13). As an example, for the
431: oscillator with $\omega=\,20meV$, $N_F(T=\,10K)=\,59$ and
432: $N_F(T=\,200K)=\,8$.
433:
434: Inspection of (14) offers the key to perform reliable path
435: integrations according to the sign of the $\phi^4$ potential. At
436: high temperature, a large contribution to $Z_{osc}$ is expected to
437: come from the paths having $u_o$ which maximizes $\exp\bigl(\kappa
438: f(u_o)\bigr)$ with $f(u_o)=\, -(a u_o + u_o^2 + b u_o^4), \, a
439: \equiv \,\beta \bar g/\kappa; \,\, b \equiv\,\beta\delta/4\kappa$.
440: In general, we find that for a hard (soft) potential, the
441: $du_o$-integration has to be carried out along the $u_o < 0 \,(u_o
442: > 0)$ axis, with cutoff $|u_o|_{max} \sim 0.6/\sqrt{\kappa}$. This
443: permits to include the set of paths which mainly contribute to the
444: euclidean oscillator action. On the Fourier coefficients
445: integrals, $\int da_n \,\int db_n$, we set the cutoffs
446: $|a_n|_{max}, \, |b_n|_{max} \sim 0.6/\sqrt{\gamma_n}$ both for
447: hard and soft potentials thus achieving numerical convergence and
448: correct computation of the Gaussian integrals once the non
449: linearities are switched off. It turns out, see the definitions in
450: (14), that the cutoffs on the oscillator path integration are
451: increasing functions of temperature ($\propto \sqrt{T}$)
452: consistently with the physical expectations of large amplitude
453: displacements at high $T$. As the path displacements encounter an
454: upper limit due to the cutoffs, $u(\tau) \leq |u_o|_{max} +
455: 2\sum_{n=1}^{N_F}|a_n|_{max}$, the distribution of on site
456: potentials has a lower limit also in the case of soft and
457: attractive non linearities. This avoids numerical divergences and
458: makes the problem physically meaningful.
459:
460:
461: \section*{V. Results}
462:
463: We test the relevance of the non linearities on the equilibrium
464: thermodynamics of the system and present the calculation for the
465: heat capacity in the intermediate to low temperature range.
466:
467: Figures 1 show the behavior of a low energy ($\omega=\,20meV$)
468: oscillator without ($g=\,0$) and with ($g=\,2,4$) coupling to the
469: electronic subsystem in the adiabatic regime: $t/ \omega =\,5$.
470: Figs.1(a),(b) assume an anharmonic potential with positive quartic
471: force constant $\delta = 10^3, 10^4 meV \AA^{-4}$ respectively.
472: Also the harmonic heat capacity is reported on for comparison. The
473: hard potential lifts the free energy over the harmonic values with
474: a more pronounced enhancement at increasing $T$ and for larger
475: $\delta$. The effect on the free energy second derivative is
476: however scarce and essentially consists, see Fig.1(a), in a slight
477: increase ( decrease) of the heat capacity at low (high)
478: temperatures. The reduction of the heat capacity (with respect to
479: the harmonic plot) at intermediate and high $T$ is more evident in
480: Fig.1(b) where the quartic force constant is larger. This result
481: is in accordance with diagrammatic perturbative treatments of the
482: anharmonic crystals which predict negative contributions to the
483: constant volume specific heat arising from positive force
484: constants in the quartic potential \cite{martin,io90}. The {\it
485: e-ph} coupling strength also tends to decrease the oscillator heat
486: capacity but the overall effect is small as the insets in the
487: figures show: infact the electronic term dominates the total heat
488: capacity $C$ and the oscillator contribution is not
489: distinguishable in the plots of the $C$ over temperature ratios
490: versus $T$. The low temperature upturn is due to the large
491: electron energy term in (9) and precisely ascribable to the
492: feature of the variable range (on the $\tau$ scale) of the
493: electron hopping, captured by the path integral formalism
494: \cite{io3}.
495:
496: The cases of a soft on site potential are reported on in Fig.1(c)
497: with $\delta = -10^3 meV \AA^{-4}$ and Fig.1(d) with $\delta =
498: -10^4 meV \AA^{-4}$. The characteristic potential parameter is
499: $M\omega^2/|\delta| \simeq 0.5$ and $0.05 \AA^2$, respectively.
500: The potential $V(u(\tau))=\, {M}\omega^2 u(\tau)^2/2 +
501: {{\delta}}u(\tau)^4/4$ is attractive for those paths such that
502: $u(\tau)^2 > 2M\omega^2/|\delta|$.
503:
504: At any $T$, we integrate over a distribution of time dependent
505: potentials. Thus, at a given lattice site, the electron may
506: experience an attractive or repulsive scattering centre according
507: to the size of the atomic path. As an example, at $T=\,200K$, we
508: get a maximum path $u_{max}$ such that $u_{max}^2 \sim 0.15
509: \AA^2$. This guarantees that $V(u(\tau))$ is generally repulsive
510: in Fig.1(c) and attractive for a broad class of paths in Fig.1(d).
511:
512: The effects on the oscillator heat capacity are twofold: i) the
513: soft potential enhances the heat capacity mainly in the low $T$
514: range with respect to the hard potential and this feature is much
515: more pronounced in Fig.1(d); ii) the trend of the {\it e-ph}
516: coupling is opposite to that observed in Figs.1(a),(b): now by
517: increasing the $g$ values one gets higher heat capacities although
518: the size of this effect is small on the scale of the electronic
519: terms in the adiabatic regime as revealed by the $C/T$ plots in
520: the insets.
521:
522: Let's come to the antiadiabatic regime ($t/ \omega =\,0.5$)
523: discussed in Figures 2, where a low harmonic energy ($\omega
524: =\,10meV$) is assumed to emphasize the size of the anharmonicity
525: together with a very narrow electron band. A hard on site
526: potential with $\delta = 10^3 meV \AA^{-4}$ is taken in Fig.2(a):
527: the shape of the oscillator heat capacity signals the effects of
528: the non linearities which flatten the curve at intermediate $T$
529: and enhance the heat capacity also at low $T$ with respect to the
530: corresponding case of Fig.1(a). As the Debye temperature is now
531: smaller (than in Fig.1(a)) the hard anharmonicity decreases the
532: constant volume heat capacity with respect to the harmonic plot
533: over a broader temperature range.
534:
535: The {\it e-ph} coupling plays a minor role also in antiadiabatic
536: conditions. The anharmonic contribution is visible in the total
537: heat capacity as the inset makes evident although the dominant
538: electronic feature persists in the low $T$ upturn of $C/T$. In
539: Fig.2(b), we assume $g =\,2$ and consider two cases of soft
540: potential: the oscillator anharmonicity becomes relevant and such
541: to modify the shape of the anomalous upturn in the total heat
542: capacity. An enhancement of the $C/T$ values is observed at
543: intermediate and low $T$ and, in the case of the largest
544: $|\delta|$ generating a soft attractive potential, the oscillator
545: heat capacity $C_{osc}$ yields an upturn in $C_{osc}/T$
546: independently of the electronic term.
547:
548: \section*{VI. Conclusions}
549:
550: We have studied the path integral of the one dimensional non
551: linear Holstein model in which a set of dispersionless oscillators
552: provides the environment for the electron. The model is
553: semiclassical as the lattice displacements are treated classically
554: while the electron operators are thermally averaged over the
555: ground state Hamiltonian. The {\it e-ph} coupling of the model is
556: local and generates a perturbing current which linearly depends on
557: the oscillator path amplitude $u(\tau)$ where $\tau$ is the time
558: (or inverse temperature) of the Matsubara Green functions
559: formalism. The anharmonicity on the lattice site is modelled
560: through a $\phi^4$ potential that may result attractive for a set
561: of displacement paths in the case of a negative quartic force
562: constant (soft potential). We have derived the path integral of
563: the interacting system and computed the total partition function
564: selecting, as a function of the temperature ($T \leq 200K$), both
565: the electron and oscillator paths which yield the largest
566: contribution to the action. While quantum electron paths have
567: increasing amplitudes at decreasing temperatures, the atomic
568: displacements are growing functions of $T$. This relevant physical
569: feature is accounted for in our model as the cutoffs on the
570: electron path integration are proportional to the electron thermal
571: wavelength whereas, on the atomic path integration, we find
572: cutoffs proportional to $\sqrt{T}$.
573:
574: The oscillator partition function includes the effect of the
575: coupling to the electron subsystem but the on site anharmonicities
576: play a major role mainly when the potential is soft and the
577: harmonic energy is low. Among the thermodynamic properties we have
578: chosen to present the heat capacity $C$ in view of the upturn in
579: the $C/T$ behavior due to the low $T$ electron hopping tuned by
580: the value of the overlap integral in the Hamiltonian model. The
581: computation is highly time consuming especially because of the
582: strong mixing of the path Fourier components generated by the non
583: linear potential.
584:
585: In general we find that: i) in the case of a hard quartic
586: potential, switching on the {\it e-ph} coupling leads to lower the
587: free energy and its second derivative, ii) when the quartic
588: potential is soft, {\it e-ph} coupling and anharmonicity act
589: sinergically enhancing the thermodynamic functions.
590:
591: These results can be understood on general physical grounds.
592: Infact, a hard quartic potential shifts the characteristic phonon
593: frequency upwards thus hardening the spectrum and broadening the
594: size of the quasiparticle. If the {\it e-ph} coupling is enhanced
595: (independently of the on site anharmonicity) the magnitude of the
596: source which causes the lattice distortion becomes larger. This
597: further hardens the vibrations and leads to decrease the
598: anharmonic heat capacity over the whole temperature range.
599: Instead, when the quartic potential is soft the phonon frequency
600: is lowered and the oscillator potential well is more flexible. In
601: this case, larger {\it e-ph} coupling strengths favour the
602: self-trapping of quasiparticles with heavier effective masses.
603: This is physically equivalent to soften the phonon spectrum and
604: enhance the heat capacity.
605:
606: The electron contribution to the heat capacity is dominant in the
607: adiabatic regime whereas antiadiabatic systems are expected to
608: present significant anharmonic corrections. Infact, in the
609: antiadiabatic regime the quasiparticle is a small size object on
610: the lattice scale and the electron energy associated with the
611: overlap integral is small. Thus, this regime proposes a physical
612: picture in which the electron hardly hops from site to site and
613: its effective mass becomes heavier on the scale of the atomic
614: mass. But a potential well generated by "lighter oscillators" is
615: more sensitive to on site anharmonic effects.
616:
617: In particular, soft potentials increase the heat capacities over
618: the harmonic values and reinforce the upturn in the $C/T$ versus
619: $T$ plots when the on site anharmonicity is such to produce
620: attractive potentials for a set of lattice displacement paths.
621: Since the path amplitudes are larger at high $T$, soft attractive
622: potentials induce rapidly increasing phonon heat capacities at
623: growing $T$ as shown in Fig.2(b).
624:
625: Thus our path integral investigation and the thermodynamical
626: results point to a complex role of the lattice anharmonicities in
627: the one dimensional Holstein model and suggest that on site
628: potentials may be experienced as attractive or repulsive according
629: to the temperature and the amplitude of the atomic path. Such
630: potentials may provide scattering centres generating local
631: disorder whose effect on the system thermodynamics is superimposed
632: to the disorder induced by the hopping of electrons with variable
633: range.
634:
635:
636: \begin{figure} \vspace*{18truecm}
637: \caption{(Color online) Anharmonic Oscillator Heat Capacities
638: versus temperature for three values of {\it e-ph} coupling $g$ and
639: oscillator energy $\omega =\,20meV$. (a) Hard potential with force
640: constant $\delta = 10^3 meV \AA^{-4}$; (b) $\delta = 10^4 meV
641: \AA^{-4}$; (c) soft potential with $\delta = -10^3 meV \AA^{-4}$;
642: (d) $\delta = -10^4 meV \AA^{-4}$. The harmonic heat capacity is
643: plotted in (a) and (b) for comparison. The insets show the {\it
644: Total (electronic plus anharmonic oscillator) Heat Capacity over
645: temperature} ratios in the adiabatic regime $t/ \omega =\,5$.}
646: \end{figure}
647:
648: \begin{figure} \vspace*{12truecm}
649: \caption{(Color online) Anharmonic Oscillator Heat Capacities
650: versus temperature in the antiadiabatic regime $t/ \omega =\,0.5$
651: and oscillator energy $\omega =\,10meV$. The harmonic heat
652: capacity is also plotted. (a) Hard potential force constant
653: $\delta = 10^3 meV \AA^{-4}$ with three values of {\it e-ph}
654: coupling $g$. (b) Two soft potential force constants at fixed {\it
655: e-ph} coupling. The insets show the {\it Total (electronic plus
656: anharmonic oscillator) Heat Capacity over temperature} ratios. The
657: electronic contribution is plotted separately for comparison. }
658: \end{figure}
659:
660:
661:
662:
663: \begin{references}
664: \bibitem{kaat}
665: G.A. Kaat, K. Flensberg, cond-mat/0411173
666: \bibitem{mah}
667: L.M.Woods, G.D.Mahan, Phys.\ Rev.\ B {\bf 61}, 10651 (2000).
668: \bibitem{han}
669: K.Hannewald, V.M.Stojanovi\'{c}, J.M.T.Schellekens, P.A.Bobbert,
670: G.Kresse, J.Hafner, Phys.\ Rev.\ B {\bf 69}, 075211 (2004).
671: \bibitem{alexan}
672: S.S. Alexandre, E. Artacho, J.M. Soler, H. Chacham, Phys.\ Rev.\
673: Lett. {\bf 91}, 108105 (2003).
674: \bibitem{sch}
675: T. Schneider, R. Khasanov, K. Conder, H. Keller, J. Phys. Condens.
676: Matter {\bf 15}, L763 (2003).
677: \bibitem{lan}
678: A. Lanzara, P.V. Bogdanov, X.J. Zhou, S.A. Kellar, D.L. Feng, E.D.
679: Lu, T. Yoshida, H. Eisaki, A. Fujimori, K. Kishio, J.-I.
680: Shimoyama, T. Noda, S. Uchida, Z. Hussain, Z.-X.Shen, Nature {\bf
681: 412}, 510 (2001).
682: \bibitem{chris}
683: Y. Zolotaryuk, P.L. Christiansen, J.J. Rasmussen, Phys.\ Rev.\ B
684: {\bf 58}, 14305 (1998).
685: \bibitem{voulga}
686: N.K. Voulgarakis, G.P. Tsironis, Phys.\ Rev.\ B {\bf 63}, 14302
687: (2001).
688: \bibitem{zieg}
689: D.Schneider, K.Ziegler, cond-mat/0504688.
690: \bibitem{pietro}
691: L.Boeri, G.B.Bachelet, E.Cappelluti, L.Pietronero, Phys.\ Rev.\ B
692: {\bf 65}, 216501 (2002).
693: \bibitem{cava}
694: T.Yildirim, O. G\"ulseren, J. W. Lynn, C. M. Brown, T. J. Udovic,
695: Q. Huang, N. Rogado, K. A. Regan, M. A. Hayward, J. S. Slusky, T.
696: He, M. K. Haas, P. Khalifah, K. Inumaru, and R. J. Cava, Phys.\
697: Rev.\ Lett. {\bf 87}, 037001 (2001).
698: \bibitem{fehi}
699: R.P.Feynman, A.R.Hibbs, {\it Quantum Mechanics and Path Integrals}
700: Mc Graw-Hill, New York (1965).
701: \bibitem{kleinert}
702: H.Kleinert, {\it Path Integrals in Quantum Mechanics, Statistics
703: and Polymer Physycs} World Scientific Publishing, Singapore
704: (1995).
705: \bibitem{eck}
706: U.Eckern, M.J.Gruber, P.Schwab, cond-mat/0503369.
707: \bibitem{io5}
708: M.Zoli, Phys.\ Rev.\ B {\bf 71}, 184308 (2005).
709: \bibitem{holst1}
710: T.Holstein, Ann.\ Phys.\ (N.Y.) {\bf 8}, 343 (1959).
711: \bibitem{raedt2}
712: H.De Raedt, A.Lagendijk, Phys.\ Rev.\ B {\bf 30}, 1671 (1984).
713: \bibitem{io98}
714: M.Zoli, Phys.\ Rev.\ B {\bf 57}, 10555 (1998).
715: \bibitem{atin}
716: M.Zoli, A.N.Das, J.Phys.: Cond. Matter {\bf 16}, 3597 (2004).
717: \bibitem{alekor}
718: A.S.Alexandrov, P.E.Kornilovitch, Phys.\ Rev.\ Lett. {\bf 82}, 807
719: (1999).
720: \bibitem{holst}
721: T.Holstein, Ann.\ Phys.\ (N.Y.) {\bf 8}, 325 (1959).
722: \bibitem{kenkre}
723: V.M.Kenkre, H.-L.Wu, I.Howard, Phys.\ Rev.\ B {\bf 51}, 15841
724: (1995).
725: \bibitem{trug}
726: C.Li-Ku, S.A.Trugman, J.Bon\v{c}a, Phys.\ Rev.\ B {\bf 65},
727: 174306 (2002).
728: \bibitem{io4}
729: M.Zoli, Phys.\ Rev.\ B {\bf 70}, 184301 (2004).
730: \bibitem{feynman}
731: R.P.Feynman, Phys. Rev. {\bf 97}, 660 (1955).
732: \bibitem{hamann}
733: D.R.Hamann, Phys.\ Rev.\ B {\bf 2}, 1373 (1970).
734: \bibitem{io3}
735: M.Zoli, Phys.\ Rev.\ B {\bf 67}, 195102 (2003).
736: \bibitem{fey}
737: R.P.Feynman, {\it Statistical Mechanics: A Set of Lectures}
738: W.A.Benjamin, Inc. (1972).
739: \bibitem{martin}
740: D.L.Martin, Can.J.Phys. {\bf 65}, 1104 (1987).
741: \bibitem{io90}
742: M.Zoli, Phys.\ Rev.\ B {\bf 41}, 7497 (1990).
743: \end{references}
744:
745: \end{document}
746: