1:
2: \documentstyle[amssymb,aps,preprint]{revtex}
3: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
4: %TCIDATA{OutputFilter=LATEX.DLL}
5: %TCIDATA{Created=Mon Jan 17 18:21:08 2000}
6: %TCIDATA{LastRevised=Fri Jul 14 14:17:17 2000}
7: %TCIDATA{<META NAME="GraphicsSave" CONTENT="32">}
8: %TCIDATA{<META NAME="DocumentShell" CONTENT="Journal Articles\REVTeX - APS and AIP Article">}
9: %TCIDATA{CSTFile=revtxtci.cst}
10:
11: \begin{document}
12: \title{Finite-wavevector Jahn-Teller-pairing and superconductivity in the cuprates. }
13: \author{D.Mihailovic and V.V.Kabanov}
14: \address{Jozef Stefan Institute, Jamova 39, 1001 Ljubljana, Slovenia}
15: \maketitle
16:
17: \begin{abstract}
18: A model interaction is proposed in which pairing is caused by a{\em \
19: non-local }Jahn-Teller (JT) -like instability due to the coupling between
20: planar O states and $k\neq 0$ phonons. Apart from pairing, the interaction
21: is found to naturally allow metallic stripe formation. The consequences of
22: the model for superconductivity in the cuprates are discussed. The model is
23: shown to be consistent with numerous sets of experimental data in quite some
24: detail.
25: \end{abstract}
26:
27: \pacs{74.20., 74.20.Mn, 74.72., 74.72.Dn}
28:
29: \newpage
30:
31: \section{Introduction}
32:
33: A Jahn-Teller (JT) polaron pairing effect was originally proposed as a
34: possible explanation for the superconductivity in La$_{2-x}$Ba$_{x}$CuO$_{4}$
35: by Bednorz and M\"{u}ller\cite{BednorzMuller}. Since then the JT effect has
36: been discussed by a number of authors in different contexts\cite
37: {Gorkov,Weger,Markiewicz,Kresin,EgamiJT} and although many features have
38: been observed experimentally supporting the general concept of JT\ polarons
39: \cite{Muller,GuoMeng}, so far no generally applicable model has been shown
40: to be compatible with the overall phenomenology observed in the cuprates.
41: One of the major problems is that the single-ion JT energy splitting between
42: Cu $d_{x^{2}-y^{2}}$ states and d$_{3r^{2}-z^{2}}$ states is thought to be
43: of the order of 1 eV or more, too large to play a role in the pseudogap
44: physics, which is believed to be the energy scale of the pairing
45: interaction, which is of the order of 0.1 eV. Nevertheless, the observation
46: of a large isotope effect on both $T_{c}$\cite{isotopeTc}, $T^{\ast }$\cite
47: {GuoMeng} and penetration depth\cite{pendep} firmly establishes a role for
48: lattice polarons in the pairing mechanism, while the fact that a depression
49: in the spin susceptibility usually appears at a lower temperature than the
50: ''pseudogap'' observed by charge excitation spectroscopies\cite{Mook,NMR}
51: suggests that a lattice pairing mechanism is primary and the spin ordering
52: follows.
53:
54: In this paper we outline a new type of microscopic pairing scenario in La$%
55: _{2-x}$Sr$_{x}$CuO$_{4}$ driven by a finite-wavevector JT instability. We
56: find that the proposed model can explain many of the general features both
57: in the underdoped and overdoped regions of the phase diagram and is
58: fundamentally compatible with the overall phenomenology of the cuprates.
59:
60: The experimental observations on which the present scenario is based are
61: mainly those showing evidence for the existence of dynamic incommensurate
62: lattice distortions associated with doped holes. Inelastic neutron
63: scattering \cite{Mook,Egami}, neutron PDF\cite{Sendyka}, EXAFS\cite
64: {Bianconi,Bozin} and ESR\cite{Kochelaev} experiments all show the existence
65: of dynamic lattice distortions on timescales relevant for pairing of 10$%
66: ^{-13}-$10$^{-15}$s. The inelastic neutron scattering data\cite{Egami,Mook}
67: can be singled out for {\em directly} giving not only the energy, but also
68: the wavevector associated with the lattice distortion and its range in $k$%
69: -space without any interpretation or modeling. The observed distorted
70: regions appear to be along the ($\zeta ,0,0$) (or ($0,\zeta ,0$))
71: directions, and have typical dimensions in real space of $2a\times 5a$ $%
72: (8\times 20$ \AA ), where $a$ is the lattice constant. A shematic diagram of
73: the distortion derived from an analysis of the data is shown in Figure 1.
74: The energy of the anomaly of $E_{a}=65\sim 85$ meV is of the order of the
75: ''pseudogap'' energy, while its width of $\Delta E\simeq 5$ meV corresponds
76: closely to the linewidth expected from the measured pair recombination rate,
77: $h/\pi c\tau \simeq 4$ meV\cite{Demsar}. At the doping level of $x=0.15$,
78: the 2D volume of the object in Figure 1 contains approximately 1.5 carriers.
79: Taken together, the implication is that the objects can be interpreted as $k$%
80: -space ''snapshots'' of individual pairs.
81:
82: Additional experimental observations which we consider important in the
83: present context is evidence for the co-existence of two carrier types in a
84: large part of the phase diagram\cite{Mullersusc,Bled}, and - in addition to
85: the ''pseudogap'' - the appearance of a temperature-dependent
86: superconducting gap $\Delta _{c}(T)$ which closes at $T_{c}$\cite
87: {Norman,Demsar} which is particularly well observed at higher doping levels
88: and has a magnitude at $T=0$ of $\Delta _{c}(0)\lesssim \Delta _{p}$.
89:
90: \section{JT\ pairs and stripes.}
91:
92: Before proceeding with the analysis of the {\it e-p} coupling for the case
93: of general $k$, let us briefly discuss the $\Gamma -$ point coupling ($k=0$)
94: in the tetragonal group $D_{4h}^{17}$ (I4/mmm) applicable to La$_{2-x}$Sr$%
95: _{x}$CuO$_{4}.$ The symmetrised cross product of the representations at the $%
96: \Gamma $ point is
97: \begin{equation}
98: \lbrack E_{u}\times E_{u}]=[E_{g}\times E_{g}]=A_{1g}+B_{1g}+B_{2g},
99: \end{equation}
100: which contains no degenerate representations. On the other hand, the lattice
101: vibrations at the $\Gamma -$point transform as:
102: \[
103: \Gamma =2A_{1g}+4A_{2u}+B_{2u}+2E_{g}+5E_{u}.
104: \]
105: Since there are no $B_{1g}$ and $B_{2g}$ representations at the $\Gamma $
106: point, electrons can couple only with $A_{1g}$ phonon modes. In $D_{4h}$
107: there are two such modes associated with apex oxygens or La ions. However,
108: the experiments show that the modes involved in the intraction are those of
109: in-plane O atoms, which do not couple at the $\Gamma $ point. This leads us
110: to the main conjecture of the proposed pairing model, namely the existence
111: of {\em intersite }pairs which form via a $k\neq 0$ interaction.
112:
113: The existence of intersite pairs in cuprate superconductors is inferred from
114: their very short coherence length. Given that the pair dimensions $l_{p}$
115: cannot exceed the coherence length, i.e. $l_{p}\lesssim \xi $, we may infer
116: that any possible lattice distortions associated with pairing have a finite
117: range $\sim l_{p}.$ The effect of such lattice distortions should also be
118: evident in reciprocal space, with an anomaly centered around a wavevector $%
119: k\simeq 1/l_{p}$. Following the inelastic neutron scattering data \cite
120: {Egami} which shows an anomaly approximately at $k_{0}\simeq \lbrack \pm \pi
121: /2a,0,0]$ extending over almost half the Brillouin zone (BZ) $\Delta k\sim
122: 1/2a,$ we can write the electron-phonon interaction for such an object in
123: the form:
124: \begin{equation}
125: g(k_{0},k)=g_{0}/((k-k_{0})^{2}+\gamma ^{2})
126: \end{equation}
127: where $g_{0}$ is a constant describing the strength of coupling, $k_{0}\,$\
128: defines the wavevector associated with the interaction and its range in $k$%
129: -space, which - neglecting fluctuations - also defines its extent in real
130: space (inter-hole spacing) as $l_{p}\sim k_{0}^{-1}$. $\gamma =\Delta k$
131: defines its width in $k$-space and gives the width of the distribution of
132: inter-carrier distances within the interacting pair. This is related to the
133: average size of the deformation of each particle in real space $\gamma ^{-1}$%
134: .
135:
136: We now proceed with an analysis of the $e-p$ coupling using group theory for
137: $k\neq 0\,\ $intersite pairing and first discuss the relevant phonon modes.
138: The BZ corresponding to the tetragonal space group $D_{4h}^{17}$ applicable
139: for La$_{2-x}$Sr$_{x}$CuO$_{4}$ is shown in Figure 2. To consider local
140: pairs and/or stripes forming along the Cu-O bond direction or along 45$%
141: ^{\circ }$ to it, we need to consider the general wavevector $\Sigma $ and
142: the $\Delta $ points, corresponding to the ($\zeta ,0,0)$ and ($\zeta ,\zeta
143: ,0)$ directions respectively. (The special symmetry points ($\Gamma ,X$ and $%
144: M$ etc.) give rise to commensurate distortions which will be discussed
145: later.) The relevant lattice deformation associated with the neutron mode at
146: 75 meV \cite{Egami} (Fig. 1) is of $\tau _{1}$ - symmetry, where $\tau _{1}$
147: is the irreducible representation of the little group corresponding to the $%
148: \Sigma $ direction in the BZ as shown in Figure 3. Since in principle all
149: modes of $\tau _{1}$ symmetry can couple to electrons, for completeness we
150: show all the possible modes with $\tau _{1}$ symmetry in Figure 3. However,
151: the most relevant mode - i.e. the one for which the anomaly is observed to
152: be most pronounced - involves in-plane O1 displacements along the Cu-O bonds
153: (see also Fig.1).
154:
155: \subsection{$k\neq 0$ phonon coupling to non-degenerate electronic states$.$}
156:
157: Since the $\Sigma $ point has a four pronged star in $D_{4h}$, the coupling
158: of electrons in single non-degenerate electronic states to $k\neq 0$ phonons
159: can be written as:
160: \begin{equation}
161: H_{int}=\sum_{{\bf l},s}n_{{\bf l},s}\sum_{k_{0}=1}^{4}\sum_{{\bf k}}g(k_{0},%
162: {\bf k})\exp {(i{\bf kl})}(b_{-{\bf k}}^{\dagger }+b_{{\bf k}})
163: \end{equation}
164: where ${\bf l}$ is the site label, and
165: \begin{equation}
166: g(k_{0},{\bf k})=g(\pi \gamma ^{2})^{1/2}/((k-k_{0})^{2}+\gamma ^{2})
167: \end{equation}
168: where $k_{0}$ are the 4 wavevectors corresponding to the prongs of the star
169: associated with the interaction. The nondegenerate electronic states in this
170: interaction allowed by symmetry are associated with $p_{z}$-orbitals of
171: planar oxygens, and transform as $A_{2u}$ or $B_{2u}$ representations of the
172: $D_{4h}$ symmetry group. However, the Hamiltonian (3) above on its own does
173: not lead to symmetry breaking, and thus is not of direct relevance for pair
174: or stripe formation.
175:
176: \subsection{$k\neq 0$ phonon coupling to degenerate electronic states
177: (Jahn-Teller-like pairing)$.$}
178:
179: A more interesting case arises when two-fold degenerate levels (for example
180: the two $E_{u}$ states corresponding to the planar O $p_{x}$ and $p_{y}$
181: orbitals or the $E_{u}$ and $E_{g}$ states of the apical O) interact with $k%
182: \not=0$ phonons. We are particularly interested in the phonons which lead to
183: symmetry breaking and allow the formation of intersite pairs or stripes. In
184: Eq. (5) we give the invariant Hamiltonian which couples degenerate
185: electronic states to phonons transforming as the $\tau _{1}$ representations
186: of the group of wave-vector $G_{k}$. Taking into account that $E_{g}$ and $%
187: E_{u}$ representations are real and Pauli matrices $\sigma _{i}$
188: corresponding to the doublet of $E_{g}$ or $E_{u}$ transform as $A_{1g}$ ($%
189: k_{x}^{2}+k_{y}^{2}$) for $\sigma _{0}=\left(
190: \begin{array}{cc}
191: 1 & 0 \\
192: 0 & 1
193: \end{array}
194: \right) $, $B_{1g}$ ($k_{x}^{2}-k_{y}^{2}$) for $\sigma _{3}=\left(
195: \begin{array}{cc}
196: 1 & 0 \\
197: 0 & -1
198: \end{array}
199: \right) $, $B_{2g}$ ($k_{x}k_{y}$) for $\sigma _{1}=\left(
200: \begin{array}{cc}
201: 0 & 1 \\
202: 1 & 0
203: \end{array}
204: \right) $, and $A_{2g}$ ($s_{z}$) \ for $\sigma _{2}=\left(
205: \begin{array}{cc}
206: 0 & -i \\
207: i & 0
208: \end{array}
209: \right) $representations respectively, an invariant Hamiltonian is given by:
210: \begin{eqnarray}
211: H_{int} &=&\sum_{{\bf l},s}\sigma _{0,{\bf l}}\sum_{k_{0}=1}^{4}\sum_{{\bf k}%
212: }g_{0}(k_{0},{\bf k})\exp {(i{\bf kl})}(b_{-{\bf k}}^{\dagger }+b_{{\bf k}})+
213: \nonumber \\
214: &&\sum_{{\bf l},s}\sigma _{3,{\bf l}}\sum_{k_{0}=1}^{4}\sum_{{\bf k}%
215: }g_{1}(k_{0},{\bf k})(k_{x}^{2}-k_{y}^{2})\exp {(i{\bf kl})}(b_{-{\bf k}%
216: }^{\dagger }+b_{{\bf k}})+ \\
217: &&\sum_{{\bf l},s}\sigma _{1,{\bf l}}\sum_{k_{0}=1}^{4}\sum_{{\bf k}%
218: }g_{2}(k_{0},{\bf k})k_{x}k_{y}\exp {(i{\bf kl})}(b_{-{\bf k}}^{\dagger }+b_{%
219: {\bf k}})+ \nonumber \\
220: &&\sum_{{\bf l},s}\sigma _{2,{\bf l}}S_{z,{\bf l}}\sum_{k_{0}=1}^{4}\sum_{%
221: {\bf k}}g_{3}(k_{0},{\bf k})k_{0}^{2}\exp {(i{\bf kl})}(b_{-{\bf k}%
222: }^{\dagger }+b_{{\bf k}}) \nonumber
223: \end{eqnarray}
224: where
225: \begin{equation}
226: g_{i}(k_{0},{\bf k})=g_{i}(\pi \gamma ^{2})^{1/2}/((k-k_{0})^{2}+\gamma ^{2})
227: \end{equation}
228:
229: The first term in (5) describes symmetric coupling and is identical to the
230: non-degenerate case (Eq. (3)). The second and third terms describe the e-p
231: interaction corresponding to the $\Sigma $ and $\Delta $ directions
232: respectively, while the last term describes the {\em coupling to spins}\cite
233: {Khomskii}.
234:
235: The proposed interaction (5) on its own results in a splitting of the
236: degenerate states, breaking the tetragonal symmetry and resulting in a local
237: orthorhombic distortion at $k_{0}$ extending over $\gamma $ in $k$-space. It
238: can therefore lead to the formation of bound intersite pairs and/or stripes
239: with no further interactions. Of course the stability and size of such a
240: distortion will be determined by the balance of short-range attraction,
241: long-range Coulomb repulsion and kinetic energy \cite{Kuzmartsev}.
242:
243: Now let us discuss the properties of the system governed by this Hamiltonian
244: (Eq. (5)). The importance of the different terms is of course to be
245: determined by experiments. For example, the large 20\% anomaly in inelastic
246: neutron scattering at the $\Sigma $ point clearly emphasizes the second ($d$%
247: -wave) term, while the absence of strong anomalies at $k=0$ de-emphasizes
248: the symmetric ($s$-wave) term and so the new ground state is expected to be
249: a pair which extends over a few unit cells along the Cu-O bond direction $%
250: (\pm \zeta ,0,0)$ or $(0,\pm \zeta ,0)$.\cite{direction} The internal
251: lattice structure within the pair is distorted, so pairing would be
252: associated with a reduced {\it local} symmetry within the pair. In other
253: words, the tetragonal or pseudo-tetragonal symmetry of the crystal is broken
254: locally by the formation of a non-local JT pair with a binding energy $%
255: E_{JT} $ given by the solution to the Hamiltonian (5). We thus associate the
256: pairing energy gap $E_{JT}$ with the experimental observation of a
257: ''pseudogap'' at $kT^{\ast }\sim \Delta _{p}=E_{JT}$\cite{NMR,Demsar}.
258:
259: To understand these finite-wavevector JT pairs in the context of the phase
260: diagram of the cuprates, we consider the effect of thermal fluctuations as
261: the temperature is reduced through $T^{\ast }$ in the underdoped phase (Fig.
262: 4). For $T>T^{\ast }$ thermal energy prevents the carriers from forming
263: pairs at all levels of doping (shown schematically in the top row in Figure
264: 5). Approaching $T^{\ast },$ JT\ pairs start to form and exist in
265: equilibrium with unbound carriers according to chemical balance at
266: thermodynamic equilibrium$\ n_{unbound}\sim \exp [-E_{JT}/k_{B}T],$ and
267: shown schematically in the lower panel of Fig. 5b). (The doping dependence
268: of $\Delta _{p}$ which is observed to approximately follow an inverse law $%
269: \Delta _{p}\sim 1/x$ \cite{Mullersusc,Demsar,Kabanov} is suggested to be a
270: result of screening as discussed by Alexandrov, Kabanov and Mott\cite{AKM},
271: and will not be discussed further here.)
272:
273: For such non-local pairs to be stable, the energy gained by the JT pairing
274: must counteract the Coulomb repulsion between two charge carriers within the
275: pair $E_{JT}\gtrsim V_{i}$. The {\it upper limit }for the Coulumb repulsion
276: between two carriers approximately one coherence length apart is given by $%
277: V_{i}=e^{2}/4\pi \varepsilon r$ $\simeq 0.15$ eV (taking $r=1/k_{0}\simeq 2$
278: nm $\lesssim \xi _{s}$ and $\varepsilon =4$\cite{Timusk}). However, since $%
279: \varepsilon (\omega )\gg 4$ in the relevant frequency range for pairing (1 -
280: 4 THz)\cite{Timusk}, the relevant value of $V_{i}$ can be significantly
281: smaller and can be easily overcome by $E_{JT}$.
282:
283: Once pre-formed bosonic pairs exist, superconductivity can occur when phase
284: fluctuations between these pairs are sufficiently reduced so that phase
285: coherence can be established between them. This can occur by Bose
286: condensation \cite{Alexandrov,Uemura,BEC} or some form of the
287: Kosterlitz-Thouless transition \cite{Pokrovsky,EK}. In both cases the
288: critical transition temperature in the underdoped region of the phase
289: diagram is given by an expression relating $T_{c}$ to the pair density $%
290: n_{p} $ \ and effective mass $m^{\ast }$:
291: \begin{equation}
292: T_{c}\simeq \hbar ^{2}n_{p}^{2/D}/(2m^{\ast })k_{B}
293: \end{equation}
294: where $D$\ is the dimensionality of the system\cite{interactions}. An
295: important issue related to whether Bose condensation occurs or another
296: mechanism is responsible for the formation of the condensate formation is
297: the number of pairs per coherence volume $V_{\xi }$. Using the experimental
298: upper limit of coherence length in La$_{1.85}$S$_{0.15}$CuO$_{4}$ at $T=0$
299: of $\xi \simeq 20$ \AA\ and assuming for the moment a uniform carrier
300: density in the Cu-O planes, then at a carrier concentration of $x=0.15$
301: there are approximately 1.5 {\em pairs per coherence volume}. Considering
302: the experimental error and uncertainties in geometrical factors involved in
303: determining $\xi $, it is clear that a crossover seems to occur near optimum
304: doping from $n_{p}<1$ per $V_{\xi }$ to $n_{p}>1$ per $V_{\xi },$ implying a
305: crossover from a Bose-condensation to overlapping-pair superconductivity
306: scenario\cite{Uemura}. Importantly, both are consistent with the
307: finite-wavevector non-local JT-pairing interaction described here. (The
308: detailed mechanism for the formation of a phase-coherent condensate is not
309: the subject of the present paper and will not be discussed further here.)
310:
311: \subsection{Stripes}
312:
313: So far, the discussion concerned a intersite pairing-JT effect with two
314: particles involved. If more than two{\it \ }particles are involved in the
315: interaction, the effect of (5) is similar and provided $k_{0}>\gamma $ (Eq.
316: 1), a JT\ distortion can occur along a stripe, for example. The internal
317: {\em lattice }structure of such stripes is defined by the JT\ lattice
318: distortion, just as for pairs. The shape of these objects is determined
319: primarily by minimisation of the Coulomb energy, and the formation of 1D
320: stripes is clearly more favourable than 2D clusters in this respect\cite
321: {Kuzmartsev}. The incommensurability of the dynamic JT distortion given by $%
322: k_{0}$ means that the number of sites in the stripe is larger than the
323: number of carriers, resulting in a partially filled ground state. The
324: electronic wavefunction inside such stripes is {\em extended}, that is, it
325: extends throughout the entire stripe, and the macroscopic transport
326: properties in the normal state are thus expected to be dominated by hopping
327: or tunneling of carries {\it between} the stripes, rather than within them.
328: The elementary excitations of such objects are expected to be Fermionic and
329: metallic in character, which makes their statistics different than for the
330: JT pairs, which are Bosons. The JT\ stripes are expected to {\em coexist }%
331: both with JT pairs {\em and} unbound particles, with their relative
332: populations determined by chemical balance and the pair binding energy $%
333: E_{JT}$ compared to the stripe formation energy. A schematic real-space
334: ''snapshot'' picture of this phase is shown in Figure 5c). Note that because
335: we have a four-pronged star for the $\Sigma -$point distortions, four
336: different types of stripes can form, each corresponding to one of the four $%
337: k_{0}$. Since the little group at the\ $\Sigma -$point does not have
338: inversion symmetry, the stripes can have a local polarisation (i.e have a
339: ferroelectric phase). This may explain the presence of a spontaneous
340: polarisation in these materials and the appearance of a pyroelectric effect
341: in La$_{2-x}$Sr$_{x}$CuO$_{4}$\cite{Pyro1} and other cuprates\cite{Pyro2}.
342:
343: A most simple and appealing possibility is that superconductivity in the
344: presence of stripes still occurs via the same pre-formed pair scenario as
345: discussed in the previous section. However, the stripes then appear to have
346: a detrimental effect on superconductivity, because they take up carriers and
347: thus reduce the number of pairs.
348:
349: \section{Overdoped regime}
350:
351: As the density of doped holes increases with increasing doping, the spacing
352: between them becomes comparable to the pair size and they start to overlap,
353: so interactions between the pairs and stripes become important, and some
354: kind of collective or cooperative effect which extends over both types of
355: objects needs to be considered.
356:
357: The Hamiltonian in Eq. (5) introduces a number of length scales (see Fig.
358: 5c)). The first is the mean distance between the charge carriers in the pair
359: (or within the stripe) $l_{p}\simeq 1/k_{0}$. The second is the length of
360: the stripes $l_{s}$ and finally, there is the length scale $l_{c}$
361: describing the characteristic distance {\em between} the pairs or stripes,
362: which is determined simply by the carrier density.
363:
364: With increasing doping, the distance between the pairs and stripes $l_{c}$
365: decreases and increased screening reduces Coulomb repulsion, which in turn
366: leads to increased stripe length $l_{s}$. At some point $l_{c}$ becomes
367: comparable to the superconducting coherence length $\xi _{s}$, and the
368: superconducting pairs become {\em proximity coupled} to the metallic stripes
369: (Fig. 5c). In other words, superconductivity in the stripes will be induced
370: below $T_{c}$ by a JT pair-gap proximity effect. Above $T_{c}$, there is no
371: proximity coupling and so clearly the superconducting order parameter must
372: be zero in the stripes. Thus it is evident that the superconducting order
373: parameter has to be $T$-dependent within the stripes. Whence an explanation
374: for the experimentally observed {\em coexistence} of a $T$-independent
375: pairing gap (''pseudogap'') $E_{JT}$ and a $T$-dependent superconducting gap
376: $\Delta _{s}(T)$ \cite{Demsar,Norman,Claeson} for which $\Delta
377: _{s}(0)\lesssim E_{JT}.$
378:
379: The proposed model suggests a simple explanation why $T_{c}$ {\em decreases}
380: in the overdoped regime. With increased doping, the stripe length $l_{s}$
381: increases leading to increased overall metallicity, while at the same time
382: {\em the number of pairs decreases}, leading to a decrease in $T_{c}$
383: according to the formula given by Eq. (8). Eventually in the metallic,
384: nonsuperconducting phase, $l_{c}\lesssim l_{p}$ and the material becomes a
385: homogeneous metal with no pairs and hence the phase no longer supports
386: high-temperature superconductivity.
387:
388: Above $T^{\ast }$ the crossover from the underdoped to the overdoped phase
389: manifests itself in a change of non-degenerate to degenerate statistics as
390: indicated by region 1 and region 2 respectively in the phase diagram in
391: Fig.4. In principle they should be distinguishable from the temperature
392: dependence of the susceptibility for example, which should be Curie-like in
393: region 1 and Pauli-like in region 2, particularly at low temperatures (see
394: Figure 4). In contrast, the crossover from region 1 to 3 (Fig. 4) is
395: governed by excitations across the pseudogap with the temperature dependence
396: of the susceptibility given by $\chi (T)\propto 1/T^{\alpha }\exp
397: [E_{JT}/kT] $ \cite{AKM,Mullersusc,NMR}.
398:
399: \section{Discussion}
400:
401: An important issue which needs to be discussed is the effective mass of the
402: non-local JT pairs coupled by $k\neq 0$ wavector phonons. Let us consider -
403: for simplicity - only the most experimentally relevant interaction for $%
404: \Sigma $ -point coupling, i.e. the ($k_{x}^{2}-k_{y}^{2})$ term of the
405: Hamiltonian Eq.(5). In this case we can apply the Lang-Firsov\cite
406: {Lang-Firsov} transformation which will give an appropriate estimate of the
407: particle mass \cite{AKR}. In that case, the effective mass renormalization
408: is exponential:
409:
410: \begin{equation}
411: \frac{m^{\ast }}{m_{0}}=\exp {(g_{eff}^{2})}
412: \end{equation}
413: where $m_{0}$ is the bare electron mass, and
414: \begin{equation}
415: g_{eff}^{2}=\frac{1}{(2\pi )^{2}}\int d^{2}kg_{k}^{2}[1-\cos {(}ka{)}].
416: \end{equation}
417: For simplicity, here the integration is carried over the Cu-O$_{2}$ plane,
418: so $k$ refers to in-plane momentum. Assuming that the main contribution
419: comes from $k\simeq k_{0}$, ignoring the effect of $\gamma $ and
420: integrating, we obtain:
421: \begin{equation}
422: g_{eff}^{2}=g^{2}k_{0}^{4}[1-\cos {(}k{_{0}}a{)}]/8\pi
423: \end{equation}
424: This formula can be rewritten in terms of the ground state energy of a
425: single polaron as:
426: \begin{equation}
427: g_{eff}^{2}=\frac{E_{p}[1-\cos {(}k{{_{0}}}a{)}]}{2\omega }
428: \end{equation}
429: where the polaron binding energy $E_{p}=g^{2}k_{0}^{4}\omega /4\pi $. When
430: compared with the similar expression for the effective mass in the Holstein
431: model which has no $k$-depedence, we find that the effective mass exponent
432: is a factor 2 smaller than the corresponding expression in the Holstein
433: (bi)polaron\cite{Bonca}. If $k_{0}<\pi /2a,$ the effective mass becomes even
434: smaller, reflecting the fact that for forward scattering the electron-phonon
435: interaction does not increase the mass strongly. Indeed for $%
436: k_{0}\rightarrow 0$, the effective mass approaches the bare electron mass $%
437: m^{\ast }\rightarrow $ $m_{0}$. This effect is similar to that discussed by
438: Alexandrov for the case of the Froehlich interaction\cite{froelich}.
439: However, note that in this case the interaction is weak and there is no pair
440: binding at all for $k=0$, which means that it is not relevant if we are
441: considering pairing, but {\it is} relevant if we consider single-electron
442: transport in the normal state. On the other hand if $k_{0}>\pi /2a,$ the
443: mass enhancement becomes more pronounced because of strong backscattering,
444: and so at the zone boundary, corresponding to the special points $X$ and $M$
445: in the BZ, we expect a very large coupling and a strongly enhanced pair
446: mass. This situation would be relevant to a zone-doubling (for the $M$%
447: -point) or quadrupling (for the $X$-point) charge density wave formation
448: and/or the formation of long-range order associated with a structural phase
449: transition. The case {\em relevant for pairing} is of course intermediate,
450: as indicated by the wavevector $k_{0}$ in the neutron experiments.
451:
452: As already discussed, according to the neutron data the interaction in the
453: cuprates appears to take place over a large range of wavevectors $\gamma $
454: centered near $k_{0}\sim 1/l_{p}$. An interesting case arises at the 1/8
455: doping level, where the interparticle distance $l=\sqrt{8}a$. If $l$
456: corresponds exactly to $l=2\pi /k_{0}$ we expect to observe a CDW\ with a
457: periodicity given by $k_{0}$. (Note that this is different to the simpler
458: case of a zone-boundary CDW\ discussed in the previous paragraph.)
459:
460: In the underdoped state the JT model is different from the bipolaronic (BP)
461: models\cite{Alexandrov,Bersukher} and other intersite models\cite
462: {Chakraverty} primarily with regard to the detailed mechanism of bipolaron
463: formation. Whereas the standard bipolaron model usually refers to
464: quantum-chemical calculations\cite{Chakraverty} and does not necessarily
465: involve a particular JT mode, nor a specific local symmetry change upon
466: pairing, the present intersite JT \ pairing model does so, and implies a
467: very specific Hamiltonian (Eq. (5)) which is based on the symmetry analysis
468: of experimentally determined local distortions. Eq. (2) at first sight has
469: some common features with the phenomenology of the charge-density wave (CDW)
470: scenario \cite{DiCastro}. The present model offers a microscopic description
471: for the origin of this interaction as arising from JT-coupling between a $%
472: k\neq 0$ mode and degenerate electronic states.
473:
474: The proposed scenario suggests the coexistence of Fermionic excitations in
475: stripes and Bosons (pairs) over the entire phase diagram in different
476: proportion determined by thermodynamic equilibrium. This appears to be born
477: out by the susceptibility data\cite{Mullersusc} and the 2-component
478: interpretation of the optical conductivity \cite{Timusk,MMM} amongst others
479: \cite{review}.
480:
481: It can also be shown to be consistent with the temperature and doping
482: dependence of angle-resolved photoemission spectra. A pairing JT deformation
483: at the $\Sigma $ point leads to objects which have finite dimensions along
484: the $a$ or $b$ crystal axes. We therefore expect to observe features
485: associated with these objects in $k$-space along the $\Sigma $ direction
486: (i.e. along $\Gamma -M)$ and the appearance of a ''pseudogap'' in\ the ARPES
487: spectra. The range of wavevectors where such a ''pseudogap'' appears is
488: given by $\Delta k\sim \gamma $ from Eq. (1). The metallic stripes on the
489: other hand, in which Fermionic excitations exist in the normal state, above $%
490: T_{c}$ we expect to observe a band which crosses the Fermi-level along the $%
491: \Sigma $ direction. Importantly, with increased carrier concentration, the
492: increased {\em coupling between pairs and stripes} leads to increased 2D
493: order, progressively extending the Fermi surface in the overdoped state.
494: Clearly, the temperature-dependent superconducting gap $\Delta _{s}(T)$
495: which forms in the stripes will appear in the same regions in $k$-space as
496: the Fermionic band. If we assume that the model can be extended to Bi$_{2}$Sr%
497: $_{2}$CaCu$_{2}$O$_{8+\delta }$, the coexistence of a $T$-dependent
498: ''superconducting'' gap and a ''pseudogap'' along $\Gamma -M$ \ (i.e. the $%
499: \Sigma $ direction : see Fig. 2), and especially the apparent ''destruction
500: of the Fermi surface'' with underdoping \cite{Norman} can be understood to
501: be consequences of the Hamiltonian (5).
502:
503: Reconciling the slight differences in the interpretation of the observed
504: lattice distortions in ESR, EXAFS and inelastic neutron scattering, the $%
505: \Sigma -$ point symmetry analysis of ionic displacements in La$_{2-x}$Sr$%
506: _{x} $CuO$_{4}$ shows that the distortion of $\tau _{7}$ symmetry at the
507: zone boundary which was invoked to explain the ESR\cite{Kochelaev} and EXAFS
508: \cite{Bianconi} (Figure 6) is in fact the zone-boundary (i.e. short-range)
509: equivalent to the $\tau _{1}$ distortion occuring over a more extended
510: length scale along the $\Sigma $ direction in the BZ and the experiments may
511: be detecting the same mode described by Eq. (5).
512:
513: We end the discussion by noting that the choice of $k_{0}$ made on the basis
514: of neutron data also determines the symmetry of the pairing channel in Eq.5.
515: The first term is isotropic ($s$-wave) while the second one has $d$-wave
516: symmetry along the Cu-O bond axes. The relative strengths of the terms are
517: of course to be determined by experiments, but the large phonon anomaly at
518: the $\Sigma $ point in the inelastic neutron data clearly emphasizes the $d$%
519: -wave component.
520:
521: \section{Conclusion}
522:
523: The main aim of the present paper is to identify an interaction which can
524: lead to pairing in La$_{2-x}$Sr$_{x}$CuO$_{4}$ on the basis of a symmetry
525: analysis of the experimentally observed anomalies in the $k\neq 0$ phonon
526: spectrum. It essentially describes the interaction which causes the
527: microscopic inhomogeneities observed in experiments. The rest of the paper
528: is devoted to a discussion of the implications for superconductivity and the
529: phase diagram. The non-local Jahn-Teller pairing interaction which couples $%
530: \tau _{1}$ modes at the $\Sigma $ point with degenerate in-plane O $p_{x}$
531: and $p_{y}$ states is in spirit, if not in detail similar to the motivation
532: described in the original paper on La$_{2-x}$Ba$_{x}$CuO$_{4}$ by Bednorz
533: and Muller\cite{BednorzMuller}. The pseudogap in the normal state results
534: from pair density fluctuations and the temperature $T^{\ast }$ represents an
535: energy scale for the pairing $kT^{\ast }\sim E_{JT}$ $\simeq 32$ meV for La$%
536: _{2-x}$Sr$_{x}$CuO$_{4}$. The model naturally leads to the formation of
537: stripes and the cross-over from a predominantly paired (Bosonic) normal
538: state to a mixed fermion-boson system in the overdoped region. A
539: straightforward and appealing way to explain the doping dependence of $T_{c}$
540: in the overdoped regime by Eq. (8) arises from the fact that at higher
541: doping levels the average stripe lengths increase and thus {\em the number
542: of pairs is reduced,} thus reducing $T_{c}$. Apart from giving rise to a
543: rather simple phase diagram which is consistent with experimental
544: observations, the model also answers the question why superconductivity
545: often appears near an orthorhombic phase of the material. However, because
546: the pairs are dynamic and incommensurate, the locally orthorhombic phase
547: associated with the JT-pair cannot be easily detected by time- and
548: spatially- averaging experimental techniques, and one does {\em not} expect
549: to observe a static orthorhombic phase below $T^{\ast }$. On the other hand,
550: the model can explain well the inelastic neutron scattering, neutron PDF,
551: EXAFS, ARPES, susceptibility and ESR, as well others\cite{Loram,Hackl,NMR}
552: which we have not discussed here.
553:
554: While here we have mainly focussed on La$_{2-x}$Sr$_{x}$CuO$_{4}$, we note
555: that similar large $k\neq 0$ lattice distortions have been reported in YBa$%
556: _{2}$Cu$_{3}$O$_{7-\delta }$\cite{YBCO} and we expect a similar mechanism to
557: work there also, as well as the other cuprates and oxides in general where
558: mesoscopic inhomogeneities are observed. We have also omitted a discussion
559: of the spin coupling associated with the local pairs given by the last term
560: in Eq. (5), but mention only that in contrast to the Holstein model, the
561: present Hamiltonian allows the formation of spin singlet {\em or }triplet
562: pairs\cite{NMR}. Finally we might add as a general comment that a short
563: superconducting coherence length of the order of the inter-carrier spacing
564: may be an indication that carriers are paired by a finite-wavevector JT
565: instability forming non-local pairs. Apart from the cuprates, alkali doped
566: fullerenes might be an example of such a case.
567:
568: \section{Acknowledgments}
569:
570: We wish to acknowledge very useful and encouraging discussions with
571: K.A.M\"{u}ller, V.Kresin, A.S.Alexandrov and T.Mertelj for important
572: comments.
573:
574: \begin{references}
575: \bibitem{BednorzMuller} G.Bednorz and K.A.M\"{u}ller, Z.Physik, {\bf B64},
576: 189 (1986)
577:
578: \bibitem{Gorkov} L.P.Gor'kov and A.V.Sokol, Pis'ma Zh.Eksp.Fiz. {\bf 46},
579: 333 (1987), JETP Lett. {\bf 46,} 420 (1987)
580:
581: \bibitem{Weger} M. Weger and R. Englman, Physica A {\bf 168}, 324 (1999)
582:
583: \bibitem{Markiewicz} R.S.Markiewicz, Physica C {\bf 200}, 65 (1992)
584:
585: \bibitem{Kresin} V.Z.Kresin A.Bill, S.A.Wolf and Yu.N. Ochinnikov, Phys.
586: Rev.B {\bf 56}, 107 (1997)
587:
588: \bibitem{EgamiJT} T.Egami, Sol.Stat.Comm. {\bf 63}, 1019 (1987)
589:
590: \bibitem{Muller} K.A.Muller, J.Superconductivity {\bf 12}, 3 (1999)
591:
592: \bibitem{GuoMeng} A.Lanzara, Guo-meng Zhao, N.L.Saini, A.Bianconi,
593: K.Conder, H.Keller and K.A.Muller J.Phys.Cond. Matter {\bf 11}, L541 (1999)
594:
595: \bibitem{isotopeTc} M.K.Crawford et al,\ Science {\bf 250},1309 (1990),
596: J.P.Franck et al, Physica C {\bf 185 - 189}, 1379 (1991), H.J.Bornemann,
597: D.Morris and H.B.Liu, Physica C {\bf 182}, 132 (1991)
598:
599: \bibitem{pendep} Guo-meng-Zhao, M.B.Hunt, H. Keller, K.A.Muller, Nature
600: {\bf 385}, 236 (1997)
601:
602: \bibitem{Mook} H.A.Mook and F.Dogan, Nature {\bf 401}, 145 (1999)
603:
604: \bibitem{NMR} D.Mihailovic, V.V.Kabanov, K.Zagar and J.Demsar, Phys.Rev.B
605: {\bf 60,} R6995 (1999)
606:
607: \bibitem{Egami} R.J.McQeeney, Y.Petrov, T.Egami, M.Yethiraj, G.Shirane and
608: Y.Endoh, Phys.Rev.Lett.{\bf 82}, 628 (1999)
609:
610: \bibitem{Sendyka} T.R.Sendyka et al, Phys.Rev.B {\bf 51,} 6747 (1995)
611:
612: \bibitem{Bianconi} A.Bianconi et al, Phys.Rev.Lett {\bf 76,} 3412 (1996),
613: N.L.Saini et al, Phys.Rev.B {\bf 55,} 12759 (1997)
614:
615: \bibitem{Bozin} E.Bozin, S.Billinge, G.H.Kwei, H.Takagi, Phys. Rev.B {\bf 59%
616: }, 4445 (1999)
617:
618: \bibitem{Kochelaev} B.I.Kochelaev et al., Phys. Rev. Lett. {\bf 79}, 4274
619: (1997)
620:
621: \bibitem{Demsar} J.Demsar, B.Podobnik, V.V.Kabanov, Th. Wolf and
622: D.Mihailovic, Phys.Rev.Lett {\bf 82}, 4918 (1999)
623:
624: \bibitem{Mullersusc} K.A.Muller, Guo-meng Zhao, K.Conder and H.Keller,
625: J.Phys:Condens. Matt. {\bf 10}, L291(1998)
626:
627: \bibitem{Bled} D.Mihailovic, I.Poberaj, T.Mertelj and J.Demsar, in
628: ''Anharmonic Properties of High-T$_{c}$ Cuprates'', Eds. D.Mihailovic et al
629: (World Scientific, 1995), p148., T.Mertelj et al Phys.Rev.B {\bf 55}, 6061
630: (1997)
631:
632: \bibitem{Norman} M.R.Norman et al, Nature {\bf 392}, 157 (1998)
633:
634: \bibitem{Khomskii} A $k$-dependent JT interaction, albeit without
635: emphasising any particular wavevector was briefly discussed in different
636: context by Kugel and Khomskii Usp.Fiz.Nauk {\bf 136} 621 (1982)
637:
638: \bibitem{Kuzmartsev} F.V.Kuzmartsev, Phys.Rev.Lett. {\bf 84}, 530 (2000)
639: and Phys.Rev.Lett. {\bf 84}, 5026 (2000), see also A.S.Alexandrov and
640: V.V.Kabanov cond-mat/0005419
641:
642: \bibitem{direction} Note that a distortion along 45$^{\circ }$ to the Cu-O
643: bonds ($\Delta -$point$)$ is also allowed, but is not discussed further
644: because the experiments suggest that the $\Sigma $-point is more relevant.
645:
646: \bibitem{Kabanov} V.V.Kabanov J.Demsar, B.Podobnik and D.Mihailovic,
647: Phys.Rev.B {\bf 59}, 1497 (1999)
648:
649: \bibitem{AKM} A.A.Alexandrov, V.V.Kabanov and N.F.Mott, Phys.Rev.Lett {\bf %
650: 77}, 4796 (1996)
651:
652: \bibitem{Timusk} D.B.Tanner and T.Timusk, in ''Physical Properties of
653: High-Temperature Superconductors III'', Ed. D.Ginsberg (World Scientific,
654: 1992)
655:
656: \bibitem{Alexandrov} A.S.Alexandrov and N.F.Mott ''{\it High-temperature
657: superconductors and other superfluids}'' (Taylor and Francis, London 1994)
658:
659: \bibitem{Uemura} Y.J. Uemura et al, Nature {\bf 364,} 605 (1993)
660:
661: \bibitem{BEC} A.S.Alexandrov and V.V.Kabanov, Phys.Rev.B {\bf 59}, 13628
662: (1999)
663:
664: \bibitem{Pokrovsky} V.L.Pokrovsky, Pis'ma Zh.Exp.Theor.Fiz. {\bf 47}, 539
665: (1988)
666:
667: \bibitem{EK} V.J. Emery and S.Kivelson, Nature {\bf 374}, 434 (1995)
668:
669: \bibitem{interactions} Of course, formula (8) is an aproximate one. For\
670: interacting pairs it needs to be modified. See for example: V.N.Popov, {\it %
671: Kontinualnie Integrali v Kvantnoi Teorii Polia i Statisticheskoi Fizike }%
672: (Atomizdat, Moscow 1976) in Russian
673:
674: \bibitem{Pyro1} I.Poberaj, D.Mihailovic, Ferroelectrics. {\bf 128}, 197
675: (1992), D. Mihailovic, I.Poberaj, Physica-C {\bf 185-189}, 781 (1991)
676:
677: \bibitem{Pyro2} D.Mihailovic, A.J.Heeger, Sol.Stat.Comm.75, 319 (1990),
678: D.Mihailovic, I.Poberaj and A. Mertelj, Phys.Rev.B. {\bf 48}, 16634 (1993)
679:
680: \bibitem{Claeson} V.M.Krasnov, A.Yurgens, D.Winkler, P.Delsing and
681: T.Claeson, Phys.Rev.Lett. {\bf 84}, 5860 (2000)
682:
683: \bibitem{Lang-Firsov} I.G.Lang and Y.A.Firsov, Zh.Exp.Teor.Fiz. {\bf 43},
684: 1843 (1963)
685:
686: \bibitem{AKR} A.S.Alexandrov, V.V.Kabanov and D.K.Ray, Phys. Rev.B, {\bf 49}%
687: , 9915 (1994)
688:
689: \bibitem{Bersukher} G.I.Bersuker, J.B.Goodenough, Physica C {\bf 274}, 267
690: (1997)
691:
692: \bibitem{Chakraverty} B.K.Chakraverty, D.D.Sarma, C.N.R.Rao, Physica C 156,
693: 413 (1988), A.S.Alexandrov, Phys.Rev B 53, 2863 (1996)
694:
695: \bibitem{DiCastro} C.Castellani, C.DiCastro, M.Grilli, Phys.Rev.Lett. 75,
696: 4650 (1995),A.Perali, C.Castellani, C.DiCastro, M.Grilli, E.Piegari and
697: A.A.Varlamov, cond-mat/9912363
698:
699: \bibitem{MMM} D.Mihailovic, T.Mertelj and K.A.M\"{u}ller, Phys. Rev. B {\bf %
700: 57}, 6116 (1998)
701:
702: \bibitem{review} For a review see for example D.Mihailovic and K.A.Muller,
703: in ''High-$T_{c}$ Superconductivity 1996: Ten Years after the Discovery'' ,
704: eds. E.Kaldis et al, \ p. 243 (Kluwer, 1997) and references therein.
705:
706: \bibitem{YBCO} Y. Petrov, T. Egami, R. J. McQueeney, M. Yethiraj, H. A.
707: Mook, F. Dogan cond-mat/0003414 (2000)
708:
709: \bibitem{Kabanov1} V.V.Kabanov and Yu.Mashtakov, Phys.Rev.B {\bf 47}, 6060
710: (1993)
711:
712: \bibitem{Bonca} Similar results were obtained by numerical exact
713: calculations with the Hubbard model, where for intersite pairs, the
714: effective mass is of the order of the single polaron mass. see J.Bonca and
715: S.Trugman (to be published, 2000)
716:
717: \bibitem{froelich} A.S.Alexandrov, Phys. Rev.Lett. {\bf 82}, 2520 (1999)
718:
719: \bibitem{Loram} J.W.Loram, K.A.Mirza, J.R.Cooper and J.L.Tallon, Physica C
720: {\bf 282-287}, 1405 (1997)
721:
722: \bibitem{Boebinger} Y.Ando et al, Physica C 282, 240 (1997), Y.Ando et al,
723: Phys.Rev. Lett. {\bf 77,} 2065 (1996)
724:
725: \bibitem{Hackl} R.Hackl, in ''The Gap Symmetry and Fluctuations in High-Tc
726: Superconductors'', Ed. J.Bok et al, (Plenum Press, N.Y. 1998), p. 249.
727:
728: \newpage
729: \end{references}
730:
731: \section{Figures}
732:
733: Figure 1. The distortion in the CuO plane corresponding to the anomalous
734: mode observed in inelastic neutron scattering in La$_{2-x}$Sr$_{x}$CuO$_{4}$
735: \cite{Egami,Mook}. The O displacements are those of the $\tau _{1}$ mode
736: shown in Figure 3 and in general have different phase.
737:
738: Figure 2. The Brillouin zone (BZ) of La$_{2-x}$Sr$_{x}$CuO$_{4}$
739: corresponding to the tetragonal phase with point group $D_{4h}.$
740:
741: Figure 3. The ionic displacements in La$_{2-x}$Sr$_{x}$CuO$_{4}$
742: corresponding to $\tau _{1}$ symmetry of the little group at the $\Sigma -$%
743: point in the BZ. The mode observed in neutron scattering corresponds to the
744: O1(1) displacements.
745:
746: Figure 4. A schematic phase diagram suggested on the basis of the proposed
747: model for the cuprates. The dashed line indicates the temperature $T^{\ast }$
748: where $kT^{\ast }\simeq E_{JT}$. The solid line indicates the temperature $%
749: T_{c}$ of the onset of macroscopic phase coherence and is given by Eq. (7).
750:
751: Figure 5. Real-space schematic diagram representing approximately 100 unit
752: cells ($\approx 4\xi ^{2}$) in the Cu-O plane at different doping levels: a)
753: for $T>T^{\ast }$ the carriers are unbound single particles (region 1 in the
754: phase diagram in Figure 4), b) for $T<T^{\ast }$in the underdoped state
755: (region 3 in Fig. 4) pairs and unbound particles co-exist with few stripes.
756: For $T<T^{\ast }$\ near optimum doping and in the overdoped state (c) and d)
757: respectively) pairs coexist with unbound particles and stripes.
758:
759: Figure 6. A superposition of two $\tau _{1}$ modes (observed in neutron
760: scattering by Egami\cite{Egami})\ with orthogonal $k$-vectors at the $\Sigma
761: $ point has the same displacements as the $\tau _{7}$ mode at the zone
762: boundary observed in ESR \cite{Kochelaev}.
763:
764: \end{document}
765: