1: \documentclass[aps,twocolumn,prl,showpacs]{revtex4}
2: %
3: \usepackage{graphicx}
4: \usepackage{amsmath}
5: \usepackage{epsfig}
6: \usepackage{dcolumn}% Align table columns on decimal point
7: \usepackage{bm}% bold math
8:
9: \begin{document}
10:
11: \title{Electron Doping of Cuprates via Interfaces with Manganites}
12:
13: \author{S. Yunoki}
14: \affiliation{
15: Department of Physics and Astronomy, The University of Tennessee,
16: Knoxville, Tennessee 37996, USA}
17: \affiliation{
18: Materials Science and Technology Division, Oak Ridge National Laboratory,
19: Oak Ridge, Tennessee 32831, USA.}
20:
21: \author{S. Okamoto}
22: \affiliation{
23: Materials Science and Technology Division, Oak Ridge National Laboratory,
24: Oak Ridge, Tennessee 32831, USA.}
25:
26: \author{S. S. Kancharla}
27: \affiliation{
28: Materials Science and Technology Division, Oak Ridge National Laboratory,
29: Oak Ridge, Tennessee 32831, USA.}
30:
31: \author{A. Moreo}
32: \affiliation{
33: Department of Physics and Astronomy, The University of Tennessee,
34: Knoxville, Tennessee 37996, USA}
35: \affiliation{
36: Materials Science and Technology Division, Oak Ridge National Laboratory,
37: Oak Ridge, Tennessee 32831, USA.}
38:
39:
40: \author{E. Dagotto}
41: \affiliation{
42: Department of Physics and Astronomy, The University of Tennessee,
43: Knoxville, Tennessee 37996, USA}
44: \affiliation{
45: Materials Science and Technology Division, Oak Ridge National Laboratory,
46: Oak Ridge, Tennessee 32831, USA.}
47:
48: \author{A. Fujimori}
49: \affiliation{Department of Physics, University of Tokyo,
50: 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan.}
51:
52: \date{\today}
53:
54:
55: %%%%%%%%%%%%%%%%%%%%%%%%
56: \begin{abstract}
57: %%%%%%%%%%%%%%%%%%%%%%%%
58: The electron doping of undoped high-$T_c$ cuprates via the transfer of charge
59: from manganites (or other oxides) using heterostructure geometries is here
60: theoretically
61: discussed. This possibility is mainly addressed via a detailed analysis
62: of photoemission and diffusion voltage experiments, which locate the Fermi level
63: of manganites above the bottom of the upper Hubbard band of some cuprate parent
64: compounds. A diagram with the relative location of Fermi levels and gaps
65: for several oxides is presented. The procedure discussed here is generic,
66: allowing for
67: the qualitative prediction of the charge flow direction at several oxide interfaces.
68: The addition of electrons to antiferromagnetic Cu oxides
69: may lead to a
70: superconducting state at the interface with minimal quenched disorder.
71: Model calculations using static and dynamical mean-field theory, supplemented
72: by a Poisson equation formalism to address charge redistribution at the interface,
73: support this view. The magnetic state of the manganites could be antiferromagnetic
74: or ferromagnetic. The former is better to induce superconductivity than the latter,
75: since the spin-polarized charge transfer will
76: be detrimental to singlet superconductivity.
77: It is concluded that in spite
78: of the robust Hubbard gaps, the electron doping of undoped cuprates
79: at interfaces appears possible, and its realization may open an exciting
80: area of research in oxide heterostructures.
81: \end{abstract}
82:
83: \pacs{73.20.-r, 74.78.Fk, 73.40.-c}
84:
85:
86: \maketitle
87:
88:
89:
90: %%%%%%%%%%%%%%%%%%%
91: \section{Introduction}
92: %%%%%%%%%%%%%%%%%%%
93:
94: The study of oxide heterostructures is rapidly developing into one of the most promising areas
95: of research in strongly correlated electronic systems. The current excitement in this
96: field was in part triggered by the recent discovery
97: of conducting interfaces, with a substantial high carrier mobility, between two insulating
98: perovskites \cite{ohtomo02,ohtomo04}.
99: These results were obtained by growing abrupt layers of the insulators LaTi(3+)O$_3$ and SrTi(4+)O$_3$.
100: When the spatial distribution of the extra electron was observed with an atomic-scale electron beam, it was
101: found to correspond to a metallic state at the interface \cite{ohtomo02}.
102: Theoretical investigations of these
103: systems \cite{okamoto04a,okamoto04b,okamoto2004},
104: using Hartree-Fock and DMFT techniques, concluded that the leakage of charge from one layer to the other explains the results.
105: Thus, charge transfer between materials in heterostructures can be used to stabilize interface states that otherwise
106: would only be obtained via the chemical doping of the parent oxide compound. This last procedure has the concomitant
107: effect of introducing quenched disorder into the sample. Then, for the preparation of doped materials without the extra complication
108: of disorder, for the realization of two-dimensional versions of perovskites,
109: and for the exploration of ``oxide electronics'' devices and functionalities,
110: the experimental and theoretical analysis of oxide heterostructures is a promising
111: and vast field, and its study is only now starting \cite{nakagawa,thiel,dagotto}.
112:
113: Before the recent developments mentioned above,
114: ferromagnetic/superconducting (FM/SC) heterostructures
115: had already received considerable attention.
116: The analysis of spin injection from
117: a ferromagnetic material to an electrode is of much importance for further progress in the area
118: of spintronics. Other properties of interest
119: in FM/SC heterostructures include: (i) spin-mixing effects
120: that can induce spin-triplet pairing correlations at the
121: interface \cite{yoshida}; (ii) Josephson couplings between two
122: singlet-superconducting layers separated by a half-metallic
123: ferromagnet \cite{eschrig}; (iii) oscillations in the singlet pairing
124: amplitude detected in SC/FM/SC geometries \cite{kontos};
125: and (iv) magnetic exchange coupling in FM/SC/FM geometries \cite{vasko,melo}.
126: Since transition metal oxides have similar lattice spacings, the combination
127: of colossal magnetoresistance (CMR) manganites \cite{dagotto,dagotto-CMR,book}
128: and high-$T_c$ cuprates
129: has been specially investigated \cite{mannhart,ahn,pavlenko}.
130: Experimental results suggest a strong FM/SC interplay resulting in
131: the injection of
132: spin polarized carriers into the SC layers \cite{vasko}.
133: Of particular importance for the purposes of
134: our investigations, the transfer
135: of charge from the La$_{1-x}$Ca$_{x}$MnO$_3$ (LCMO) ferromagnet to the YBa$_2$Cu$_3$O$_y$ (YBCO)
136: superconductor has been
137: observed in recent experiments using EELS techniques \cite{varela}. In this case, the addition of electrons, namely the
138: suppression of holes, in the YBCO component led to the concomitant suppression of superconductivity and a transition to
139: an insulating state. Similar effects
140: are also believed to occur in grain boundaries of superconductors \cite{hilgenkamp}. In other
141: LCMO/YBCO superlattice setups,
142: the superconducting critical temperature was also found to be suppressed \cite{sefrioui}.
143: This may be caused by the influence of spin-polarized
144: carriers moving from LCMO to YBCO,
145: breaking singlet Cooper pairs, or by the reduction of the number of hole carriers \cite{varela},
146: or both.
147:
148:
149:
150: It is the main purpose of the investigation reported here
151: to analyze whether the previously observed electronic charge transfer at interfaces from
152: manganites, such as LCMO or La$_{1-x}$Sr$_{x}$MnO$_3$
153: (LSMO), to hole-doped cuprates, such as YBCO,
154: can be extended to the case where the cuprate is in an undoped state. If this still occurs,
155: namely if estimations of work
156: functions lead us to believe that the Fermi level of the manganite remains above the first empty states of
157: the undoped cuprate, then this
158: Cu oxide will receive an excess of electrons, leading to the $electron$ $doping$ of an antiferromagnetic
159: (AF) state at heterostructure interfaces \cite{oh}. This
160: mixing of cuprates and manganites could lead either
161: to (i) a metallic spin-polarized state at the interface, if the manganite is ferromagnetic,
162: or to (ii) an electron-doped superconductor at low
163: enough temperatures, if the spin polarization is not strong enough to destroy Cooper pairs or if the manganite
164: used is not ferromagnetic.
165: This mechanism would allow for the electronic doping of a parent cuprate compound
166: without the complication
167: of the Coulombic and structural disorder introduced by chemical doping. The fact
168: that this is electron doping, as opposed to the already investigated interfacial hole doping, is particularly
169: interesting since a variety of material issues has prevented the electron-doped branch
170: of high-$T_c$ superconductors to develop as vastly as those
171: that are hole doped.
172:
173: Once again, note that if the manganite is ferromagnetic, then the charge transfer to the
174: undoped cuprate would occur via polarized electrons, which
175: suppress spin singlet
176: superconductivity. Thus, there are two $competing$ tendencies: adding electrons
177: favors superconductivity, but its
178: spin polarization suppresses it. The outcome is difficult to predict. Recent investigations
179: using realistic microscopic models focusing on a
180: manganite/insulator interface predict that a complex pattern of phase separation between FM and AF
181: regions could occur \cite{brey}.
182: Thus, even if superconductivity is not induced, the resulting state would be interesting,
183: due to the interplay between ferro and antiferromagnetic tendencies.
184: More importantly, note that
185: the manganite involved in the heterostructure does $not$ need to be ferromagnetic: low-bandwidth Mn oxides
186: present a wide variety of spin/charge/orbital order
187: states, other than ferromagnetism \cite{dagotto-CMR,book}. Even large-bandwidth manganites have
188: non-FM states at small and large enough hole concentration. Thus, inducing a SC
189: state is a real possibility
190: with the proper choice of the manganite partner to the cuprate.
191:
192:
193:
194: A qualitative representation of the transfer of
195: charge is shown in Fig.~\ref{systems}
196: for the cases of a doped cuprate, such as discussed in \cite{varela}, and for an undoped cuprate,
197: the focus of our investigation reported below. In case (a), SC is destroyed, while in the proposed case (b), SC is induced (if the
198: manganite is not ferromagnetic).
199: Our main result is that the novel scenario (b) appears possible. Our conclusion is based
200: on (i) a detailed analysis of the work functions of the materials
201: of relevance for these investigations, mainly involving the study of photoemission and diffusion voltage published data, as well
202: as (ii) rough
203: calculations at the mean-field and dynamical mean-field levels, showing
204: that there is no fundamental problem preventing superconductivity from happening at the
205: interface. However, it must be clearly expressed that our contribution should be considered as only the
206: first theoretical steps toward realizing electron doping at cuprate interfaces,
207: and considerable
208: more work remains to be carried out. For instance, issues such as lattice
209: reconstructions and polarity of the involved interfaces have $not$ been analyzed here (only electronic reconstructions
210: at idealized perfect interfaces were investigated). Vacancies and the most optimal location of oxygens at the interface
211: have not been investigated as well.
212: However, even with these caveats we believe the conclusions described below are sufficiently
213: interesting that they deserve experimental efforts to attempt their realization in real heterostructures.
214:
215:
216:
217: \begin{figure}[hbt]
218: \includegraphics[clip=true,width=7.5cm,angle=-0]{system.eps}
219: \begin{center}
220: \caption{
221: Schematic representation of the systems studied here.
222: (a) is a heterostructure involving a ferromagnetic
223: (FM) half-metal system, such as LCMO, and a superconductor (SC), such as YBCO. In this case, an antiferromagnetic
224: (AF) insulating state is induced at the interface. (b) is a heterostructure involving a non-FM manganite
225: and an undoped AF cuprate. Here, the transfer of charge may lead to a superconducting state at the
226: interface. Arrows indicate the direction of flow of electrons, which are accumulated
227: in the gray region.
228: }
229: \label{systems}
230: \end{center}
231: \end{figure}
232:
233:
234: The organization of the manuscript is as follows: first, we carry out a detailed analysis
235: of available experimental information to judge if the Fermi level of manganites is above
236: the lowest-energy empty state of several undoped cuprates. This is described in detail, because
237: it provides a systematic procedure to address other oxide heterostructures in the future.
238: Second, a model calculation in the mean-field approximation is carried out, both for
239: static and dynamical cases. No fundamental problem is found with the proposal of having
240: superconductivity induced at the manganite/cuprate interface. The paper concludes
241: with a discussion and summary.
242:
243:
244:
245:
246: \section{Estimation of Oxides Chemical Potential Differences
247: using Photoemission and Diffusion Voltage Experiments}
248:
249: \subsection{Overview}
250:
251: For the success of the emergent field of oxide heterostructures it is crucial to properly
252: determine the relative work functions of the materials involved, since these work functions control
253: the curvature of the valence and conduction bands (VB and CB) of the constituent materials,
254: and ultimately the carrier concentration at the interfaces.
255: Work functions of conventional metals and semiconductors have been studied for decades
256: establishing the fundamental background for current electronics.
257: Thus, for the next-generation
258: electronic devices utilizing the complex properties of correlated-electron systems,
259: such as high-$T_c$ cuprates and CMR manganites, determining the
260: relative work functions of a variety of transition-metal oxides is equally important.
261:
262: Photoemission spectroscopy (PES) is an important technique in this context.
263: PES has provided fundamental
264: information to uncover the properties of complex oxides.
265: However, although PES techniques can be of considerable help for oxide heterostructures,
266: only a limited number of experiments have been reported specifically addressing
267: the work function of transition-metal oxides \cite{Schulte01}.
268:
269: By measuring the diffusion voltage $V_d$ (or built-in potential) of a junction between two materials of
270: interest, one can also extract the chemical potential differences between them.
271: This diffusion voltage is the potential barrier at the interface
272: after the rearrangement of charge occurs.
273: If there are no extra contributions to $V_d$, such as those caused by
274: interface polarities, impurities, or lattice reconstructions, then
275: $V_d$ is equivalent to the work function difference between the two constituents.
276: If the work function of one of the materials is known, the work function of the other can be estimated.
277: This provides another procedure to study the band alignment of oxides.
278:
279: The purpose of this section is to provide a rough estimation of the band diagram of
280: perovskite transition-metal oxides (cuprates and manganites) using
281: the experimental data currently available. This is achieved by
282: combining information from chemical potential shifts obtained
283: using PES with diffusion voltage measurements on heterostructures.
284: %\newline
285:
286: \subsection{Parent compounds of High-$T_c$ cuprates}
287:
288: Since the discovery of high-$T_c$ superconductivity in the cuprates,
289: considerable experimental information has been accumulated about these compounds.
290: %
291: In particular, the chemical-potential jump from the
292: hole-doped to the electron-doped materials has been one of the important topics since
293: it is closely related to the Mott/Hubbard (or charge transfer) gap.
294: Here, let us first consider La$_2$CuO$_4$ (LCO) and Nd$_2$CuO$_4$ (NCO)
295: as the parent compounds of hole-doped
296: and electron-doped cuprates, respectively.
297: It should be mentioned that these compounds have different structures:
298: LCO has the layered perovskite or $n=1$ Ruddlesden-Popper structure,
299: while NCO presents the so-called $T'$-structure without apical oxygen.
300: Even though the electronic properties of the
301: CuO$_2$ planes are quite similar for LCO and NCO, the different overall structures
302: suggest that small differences may exist in their chemical potentials, as discussed below.
303:
304: From the optical measurements, it has been suggested that the chemical potential of
305: LCO is located 0.4~eV above the top of the valence band \cite{Suzuki89}.
306: Similar results were obtained
307: with PES techniques \cite{Ino00}. Polaronic effects are believed to be the cause of this
308: somewhat ``mysterious'' shift between the top of the valence band and the actual location
309: of the chemical potential in LCO \cite{polarons}.
310: Similarly,
311: resonant-photoemission studies on NCO revealed that the chemical potential of NCO
312: is located 0.7~eV above the valence band \cite{Allen90,Namatame90}.
313: If the positions of the valence bands are identical between LCO and NCO,
314: then their chemical potential jump becomes $0.7-0.4=0.3$~eV \cite{alternative}.
315: Although an accurate determination of the individual chemical
316: potentials of these materials remains to be done,
317: estimations based on the shifts of the O 1$s$ and Cu 2$p$ core levels indicate that
318: the chemical potential jump is at most 0.5~eV \cite{Fujimori02}. Then,
319: the positions of the valence bands can be assumed to be the same, although with a $\alt 0.2$~eV
320: uncertainty.
321:
322:
323:
324: To complete the band diagram, information about the location of
325: the unoccupied conduction, or upper Hubbard band, is needed.
326: From optical spectroscopy, it was first suggested that the
327: Mott gaps of LCO and NCO are about 2~eV and 1.5~eV, respectively \cite{Tajima89}.
328: However, numerical studies of the
329: Hubbard model on finite-size clusters using appropriate hopping amplitudes
330: for the cuprates revealed that the Mott gap in Cu oxides
331: was in fact an indirect one \cite{Tsutsui99}:
332: the top of the lower-Hubbard band is at momenta
333: $(\pm \pi/2,\pm \pi/2)$ while the bottom of the upper-Hubbard band lies at $(\pi,0)$ and $(0,\pi)$.
334: Thus, the Mott gap is indirect and smaller than the optical gap.
335: The indirect nature of this gap has been confirmed by a recent resonant x-ray study \cite{Hasan00}.
336: Considering the ratio between the magnitudes of direct gap and indirect gap from the
337: theoretical considerations and the optical gap obtained by the experiments,
338: the separation between the VB and CB of LCO and NCO is estimated to be about 1.5~eV and 1~eV, respectively \cite{BIS98}.
339: Combining all this information together,
340: the schematic band diagram for these materials can be constructed and it is shown in Fig.~\ref{fig:NCO_LCO}.
341:
342:
343:
344:
345:
346: \begin{figure}[tbp]
347: \includegraphics[width=0.55\columnwidth,clip]{NCO_LCO.eps}
348: \caption{Schematic band diagrams of NCO and LCO \cite{Fujimori02}
349: based on the chemical potential jump between NCO and LCO found in a resonant-photoemission study \cite{Namatame90},
350: revised by considering the recent discovery of an indirect Mott gap in the cuprates \cite{Hasan00}.
351: Chemical potentials (dashed lines) of NCO and LCO are located about 0.7~eV and 0.4~eV above the
352: valence bands, respectively.
353: Note that the top of the valence bands of NCO and LCO do not necessarily match, but estimations
354: discussed in the text locate them very close to one another.
355: There remains an ambiguity in the chemical potential jump between these two materials,
356: but it is believed to be at most 0.5~eV \cite{Fujimori02}. For more details see text.
357: }
358: \label{fig:NCO_LCO}
359: \end{figure}
360:
361:
362:
363:
364:
365: Next, let us address another series of high-$T_c$ cuprates: YBCO.
366: The diffusion voltage of the heterostructure between YBCO (oxygen contents unclear)
367: and 0.05wt\% Nb-doped SrTiO$_3$ (STO) (Nb$_{0.05}$-STO)
368: has been measured \cite{Muraoka04}.
369: The chemical potential of YBCO was found to be 1.5~eV below that of Nb$_{0.05}$-STO.
370:
371: To establish
372: a connection with the single-layer cuprates, note that
373: similar experiments on the diffusion voltage of the single-layer parent compound Sm$_2$CuO$_4$ (SCO) and
374: 0.01wt\% Nb-doped SrTiO$_3$ (Nb$_{0.01}$-STO) have been very recently performed \cite{Nakamura07}.
375: SCO is an isostructural material of NCO, with the same $T'$-structure and formal valences, therefore
376: NCO is expected to have a chemical potential very similar to that of SCO.
377: The recent experiments of Nakamura {\it et al.}
378: revealed that the chemical potential of SCO is 1.3~eV below that of Nb$_{0.01}$-STO \cite{Nakamura07}.
379: Although there is a slight difference in the doping
380: concentration of Nb used for the two experiments (0.05wt\% Nb for YBCO and 0.01wt\% for SCO),
381: by reducing the Nb concentration to 0.01wt\% the diffusion voltage is expected to increase
382: only by 0.1~eV or less because of band-gap narrowing effects \cite{SawaPC}.
383: Thus, by combining the diffusion voltages on SCO- and YBCO-based heterostructures
384: involving a Nb-doped STO substrate,
385: the chemical potential of SCO is estimated to be $\sim 1.5+0.1-1.3 = 0.3$~eV above that of YBCO.
386:
387: Furthermore, assuming that the behavior of SCO and NCO are identical, based
388: on the similarity of their structures, then the
389: chemical potential of LCO can be estimated by considering the chemical potential jump between NCO and LCO.
390: Figure~\ref{fig:LCO_SCO_NbSTO_YBCO_LSMO} summarizes
391: the band diagram for the various high-$T_c$ cuprates studied here.
392: From these results,
393: it is deduced that the parent compounds of hole-doped cuprates have fairly similar chemical potentials
394: despite their different crystal structures.
395: %\newline
396:
397:
398:
399:
400: \begin{figure*}[tbp]
401: \includegraphics[width=1.5\columnwidth,clip]{LCO_SCO_NbSTO_YBCO_LSMO.eps}
402: \caption{Schematic band diagrams of LCO, SCO(NCO), Nb$_{0.01}$-STO, Nb$_{0.05}$-STO, YBCO, and LSMO
403: based on diffusion voltage measurements \cite{Muraoka04,Muramatsu05,Nakamura07,SawaPC}
404: and photoemission spectroscopy \cite{Namatame90,Fujimori02,Matsuno02}.
405: Tops of valence bands (VB) and bottoms of conduction bands (CB) are indicated by solid lines,
406: while chemical potentials are indicated by dashed lines.}
407: \label{fig:LCO_SCO_NbSTO_YBCO_LSMO}
408: \end{figure*}
409:
410:
411: \subsection{ Doping dependence of the chemical potential in High-$T_c$ cuprates}
412: %
413: The doping dependence of the chemical potential is important when using doped compounds for the
414: oxide heterostructures.
415: The chemical potential shifts from their parent compounds
416: of various high-$T_c$ cuprates have
417: been intensively studied \cite{Namatame90,Ino97,Harima03,Yagi06}.
418: For the benefit of the readers,
419: the main results are summarized
420: in Fig.~\ref{fig:WorkFunctions} (a) for hole-doped compounds,
421: and Fig.~\ref{fig:WorkFunctions} (b) for electron-doped compounds \cite{Harima01}.
422:
423:
424:
425: \begin{figure*}[tbp]
426: \vskip 1.0cm
427: \includegraphics[width=1.75\columnwidth,clip]{WorkFunctions.eps}
428: \caption{Chemical-potential shifts of various transition-metal oxides, reproduced
429: from photoemission experiments.
430: (a) are the results for hole-doped cuprates.
431: Na-CCOC: Ca$_{2-x}$Na$_x$CuO$_2$Cl$_2$ \cite{Yagi06},
432: LSCO: La$_{2-x}$Sr$_x$CuO$_4$ \cite{Ino97},
433: and Bi2212: Bi$_2$Sr$_2$Ca$_{1-x}$R$_x$Cu$_2$O$_{8+y}$ (R=Pr, Er) \cite{Harima03}.
434: The absolute values at the undoped origin of each series are assumed to be the same.
435: Reproduced from \cite{Yagi06}.
436: (b) Chemical potential shift of electron-doped cuprates NCCO:
437: Nd$_{2-x}$Ce$_x$Cu$_2$O$_4$.
438: Reproduced from \cite{Harima01}.
439: (c) Chemical-potential shift of LSMO.
440: These shifts are measured at liquid-nitrogen temperature except for LaMnO$_3$. Reproduced
441: from \cite{Matsuno02}.
442: }
443: \label{fig:WorkFunctions}
444: \end{figure*}
445:
446:
447:
448:
449: There are still considerable discussions
450: on the electronic properties of the underdoped region of La$_{2-x}$Sr$_x$CuO$_4$ (LSCO), involving concepts such as
451: stripe formation, phase separation, and spin glass.
452: Those may explain the flat chemical potential shifts of LSCO at $0\le x < 0.15$.
453: However, focusing on the overdoped region where the mixed-phase complexity is reduced, then the
454: chemical potential shifts behave linearly with hole doping
455: with a slope of about 0.2 eV per 0.1 carrier concentration, which is close to that of
456: electron-doped Nd$_{2-x}$Ce$_x$CuO$_4$ (NCCO).
457: This may indicate that the band structures of the CuO$_2$ planes in those compounds are similar, as widely believed.
458: On the other hand, the
459: $\Delta \mu$ vs. carrier concentration slope for Bi$_2$Sr$_2$CaCu$_2$O$_{8+y}$ (Bi2212) is larger than
460: for the single-layer compounds, perhaps due to its bilayered nature.
461:
462:
463:
464:
465:
466:
467:
468:
469: \subsection{ Connecting the High-$T_c$ cuprates with the CMR manganites}
470: %
471: As already discussed, both from the fundamental physics as well as the device engineering perspective,
472: heterostructures involving high-$T_c$ cuprates and CMR manganites are important.
473: Here, we consider the cubic perovskite manganites first, and then turn to the double-layered manganites.
474:
475: It should be mentioned that the $direct$ measurement of the work function
476: of cubic manganites has been performed by several groups
477: using PES \cite{Jong03,Minohara07} and the Kelvin method \cite{Reagor04},
478: and results are approximately consistent with one another indicating an intrinsic work function
479: $\sim 4.8$~eV for La$_{1-x}$Sr$_x$MnO$_3$ with $x=0.3,0.4$.
480: These experimental data are important in considering interfaces with the cuprates.
481: However, as the individual
482: work functions of many cuprates are not yet available but only their differences, then we
483: have to consult other experiments in order to establish a connection between
484: manganites and cuprates.
485: Here, we consider PES experiments that measured the chemical potential shift in doped manganites \cite{Matsuno02},
486: and those that provided the diffusion-voltage
487: on La$_{0.8}$Sr$_{0.2}$MnO$_3$ (20\% hole-doped LSMO)/Nb$_{0.05}$-STO heterostructures \cite{Muramatsu05}.
488:
489: According to the experiment by Matsuno {\it et al}. \cite{Matsuno02}, the
490: chemical potential of La$_{1-x}$Sr$_x$MnO$_3$ (LSMO) changes rather monotonically from the parent compound LaMnO$_3$ to
491: a value $\Delta \mu$=$-0.5$~eV for La$_{0.4}$Sr$_{0.6}$MnO$_3$ (60\% hole-doped LSMO)
492: (see Fig.~\ref{fig:WorkFunctions} (c)).
493: Furthermore, diffusion voltage measurements
494: on a La$_{0.8}$Sr$_{0.2}$MnO$_3$ (20\% hole-doped LSMO)/Nb$_{0.05}$-STO heterostructure,
495: done by Muramatsu {\it et al.} \cite{Muramatsu05}, reported a value $V_d = 0.65$~eV.
496: Combining these experiments, the
497: bands of LSMO can be aligned relative to those of the cuprates obtained in the previous section.
498: Figure \ref{fig:LCO_SCO_NbSTO_YBCO_LSMO} summarizes the band alignment of cuprates and manganites,
499: which is our main result of this section.
500: %
501: In particular, the chemical potential shifts between the various cuprates of potential
502: relevance and the 20\% doped LSMO are summarized as follows:
503: (i) for SCO $\Delta \mu = \mu_{LSMO} - \mu_{SCO}\sim 0.55$~eV,
504: (ii) for LCO $\Delta \mu = \mu_{LSMO} - \mu_{LCO}\sim 0.85$--$1.05$~eV,
505: and (iii) for YBCO $\Delta \mu = \mu_{LSMO} - \mu_{YBCO}\sim 0.85$~eV.
506: In particular, it is natural to expect that electrons be
507: transfered from LSMO to the valence band of hole-doped YBCO at an interface between the two materials.
508: This has already been observed experimentally \cite{varela}, thus it is reassuring
509: that the simple procedure followed
510: here is consistent with available experimental information \cite{YBCO-work}.
511:
512: Consider now another class of Mn oxides, the
513: double-layered manganites with the so-called $n=2$ Ruddlesden-Popper structure.
514: Direct measurements of work function using photoemission have actually been performed on
515: La$_{1.2}$Sr$_{1.8}$Mn$_2$O$_7$.
516: It has been reported that the work function of La$_{1.2}$Sr$_{1.8}$Mn$_2$O$_7$
517: increases with decreasing temperature across the Curie temperature (125~K)
518: from $\sim$~3.5~eV at 180~K to $\sim$~3.56~eV at 60~K \cite{Schulte01},
519: while the simple double-exchange model predicts the opposite trend.
520: This fact may indicate the importance of several ingredients neglected in the double-exchange model,
521: such as electron-electron interaction, electron-lattice interaction, and orbital degeneracy.
522: The important point worth emphasizing here is that
523: the work function of the double-layered manganite
524: is more than 1~eV smaller than that of cubic LSMO
525: (whose work function is $\sim4.8$~eV) \cite{Jong03,Minohara07,Reagor04},
526: which means that the chemical potential of the
527: double-layer manganite is more than 1~eV $higher$ than the cubic perovskite LSMO.
528: Thus, at least naively,
529: it should be possible to inject electrons to the undoped cuprates using the double-layer manganites.
530: However, there are issues that remain to be clarified for this conclusion to be valid, particularly
531: the role of the surface condition:
532: for double-layered manganites, the cleaved surface used in PES experiments
533: is always AO$^{2-}$ (A: A-site ion, such as La$^{3+}$ and Sr$^{2+}$),
534: thus positively charged.
535: The low work function of
536: bilayered manganites could be caused by this charged surface.
537: On the other hand, for cubic manganites the surface can be either positively charged (AO termination)
538: or negatively charged (MnO$_2$ termination).
539: For this purpose, photoemission experiments on cubic manganites with controlled surface conditions
540: are highly desirable.
541:
542: Summarizing, in this subsection we discussed the possible electron doping of cuprates from manganites at ideal interfaces.
543: In the paragraphs above, it was concluded that
544: the parent compounds of electron-doped cuprates, such as NCO and SCO, are the best candidates for this purpose
545: since they have the lowest conduction band
546: (the CB in LCO is about 0.5~eV higher than that in NCO,
547: and the CB in YBCO is expected to be even higher than those of NCO and LCO).
548: To dope electrons into a cuprate parent compound,
549: the minimal condition is that the chemical potential of manganites be higher
550: than the bottom of the conduction band (or the chemical potential
551: if they are not the same) of the undoped cuprate involved.
552: This condition seems to be satisfied for cubic manganites LSMO
553: in a substantial doping range $x$ that includes FM and AF states, when mixed with NCO or SCO \cite{STO-comment}.
554: Possible improvements may be achieved by using double-layered manganites,
555: which have about 1~eV higher chemical potential than cubic manganites.
556:
557:
558:
559:
560:
561: \section{Superconductivity at the Manganite/Cuprate Interface: Static Mean-Field Approximation}
562:
563: The simple ideas described in the previous sections on
564: the possibility of electron doping of cuprates at interfaces
565: need to be checked in more theoretical detail to confirm their
566: consistency. For this purpose, first a static mean-field study will
567: be here discussed, followed in the next section by a more detailed
568: analysis including dynamical effects. The emphasis will be given
569: to a manganite/cuprate interface, but the generation of superconductivity
570: by electron doping can occur for any other combination where
571: electrons are donated to the cuprates.
572:
573: The model used here is defined by the following
574: Hamiltonian on a three-dimensional cubic lattice:
575: \begin{equation}\label{model}
576: H=-\sum_{{\bf i},{\bf j}}\sum_s t_{\bf ij}
577: c_{{\bf i}s}^\dag c_{{\bf j}s} + H_I
578: +\sum_{\bf i}\left(\phi_{\bf i}-\mu+W_{\rm L/R} \right)n_{\bf i}.
579: \end{equation}
580: Here $c_{{\bf i}s}^\dag$ is the electron creation operator at site
581: ${\bf i}=(i_x,i_y,i_z)$ with spin $s=\uparrow,\downarrow$, $n_{\bf i}$
582: is the electron density at site {\bf i}, and
583: $t_{\bf ij}$ is the nearest-neighbor hopping, which is $t$ on the
584: $xy$ plane and $t_z$ in the $z$ direction. $\phi_{\bf i}$ is the electronic
585: potential (discussed in more detail later) that will take into account effects
586: related to the charge redistribution. $\mu$ is the chemical potential, and $W_{\rm L}$
587: ($W_{\rm R}$) are site potentials
588: for the left (right) side of the system to regulate the transfer of charge.
589: $H_I$ contains the interaction
590: terms. On the left side of the system (the manganite) the interaction is
591: \begin{equation}\label{dx}
592: H_I^{\rm (L)} = -J_{\rm H}\sum_{\bf i}\sum_{\alpha,\beta}
593: c_{{\bf i}\alpha}^{\dag}\left(\vec{\sigma}\right)_{\alpha\beta}
594: c_{{\bf i}\beta}\cdot {\vec{S}}_{\bf i},
595: \end{equation}
596: which is the standard ``double-exchange'' term.
597: $\vec{\sigma}=(\sigma_x,\sigma_y,\sigma_z)$ are Pauli matrices, and
598: ${\vec{S}}_{\bf i}$ is a classical localized spin at site {\bf i}
599: ($|{\vec{S}}_{\bf i}|=1$) representing the $t_{\rm 2g}$ spins.
600: On the right side of the system, the cuprate,
601: the standard repulsive Hubbard interaction ($U$$>$0) is
602: supplemented by
603: a nearest-neighbors attraction ($V$$<$0) that favors superconductivity
604: in the $d$-wave channel:
605: \begin{equation}\label{uv}
606: H_I^{\rm (R)} = U\sum_{\bf i}n_{{\bf i}\uparrow}n_{{\bf i}\downarrow}
607: + V \sum_{\langle{\bf i}{\bf j}\rangle}n_{\bf i}n_{\bf j}.
608: \end{equation}
609: The number operator is
610: $n_{{\bf i}s}=c_{{\bf i}s}^{\dag}c_{{\bf i}s}$,
611: and $\langle{\bf i}{\bf j}\rangle$ indicates
612: a pair of nearest-neighbor sites {\bf i} and {\bf j} on the $xy$ plane.
613: It is well known that this $t$-$U$-$V$ model leads to a rich phase diagram,
614: that includes superconductivity \cite{tuv}.
615: Periodic (open) boundary conditions are used in the
616: $x$ and $y$ ($z$) directions
617: for a lattice of size $L=L_x\times L_y\times L_z$ sites. The focus of the
618: results described below is on $L_x$=$L_y$=16, but other sizes have been
619: checked confirming that size effects are not strong in this study.
620:
621:
622: Two comments are in order: (1) A fully realistic model for manganites
623: should contain two $e_g$-orbitals and
624: a Jahn-Teller electron-phonon coupling. However, it is common practice in this context
625: to use just one
626: $e_g$-orbital for simplicity, since several phenomena
627: are common to both one and two
628: orbitals \cite{dagotto-CMR}. Moreover, the focus here is not
629: on the CMR regime but only on
630: combining a standard homogeneous manganite state with a cuprate.
631: Thus, the electron-phonon (e-ph) interaction, crucial to understand the CMR
632: effect, is here neglected for simplicity. In fact, LSMO does not have a large
633: CMR effect, thus the e-ph coupling may not be strong in this material.
634: (2) More severe is the approximation carried out on the cuprate side
635: since a mean-field approximation with an explicitly attractive force is employed.
636: However, this approximation will be relaxed below and the Hubbard
637: model will be used without an explicit attraction,
638: at the price of having to constraint the study to smaller
639: systems than those that can be analyzed in the static mean-field approximation.
640:
641:
642:
643:
644:
645:
646: To treat the many-body interactions described by $H_I^{\rm (R)}$, here a simple mean-field
647: approximation is adopted. For the first interaction term, the following replacement is introduced,
648: \begin{equation}
649: U\sum_{\bf i}n_{{\bf i}\uparrow}n_{{\bf i}\downarrow} \to
650: U\sum_{\bf i} \left [
651: \langle n_{{\bf i}\uparrow}\rangle n_{{\bf i}\downarrow}
652: + n_{{\bf i}\uparrow} \langle n_{{\bf i}\downarrow}\rangle
653: - \langle n_{{\bf i}\uparrow}\rangle \langle n_{{\bf i}\downarrow}\rangle
654: \right]. \nonumber
655: \end{equation}
656: For the second term, a standard BCS model approximation is used:
657: \begin{equation}
658: \begin{array}{l}
659: \displaystyle{
660: V \sum_{\langle{\bf i}{\bf j}\rangle}n_{\bf i}n_{\bf j} \to}\\
661: \displaystyle{
662: V \sum_{\langle{\bf i}{\bf j}\rangle}\left[\left(
663: \Delta_{\bf ij}c_{{\bf i}\uparrow}^{\dag}c_{{\bf j}\downarrow}^{\dag}
664: + \Delta_{\bf ji}c_{{\bf j}\uparrow}^{\dag}c_{{\bf i}\downarrow}^{\dag}
665: \right) + {\rm H.c.} +|\Delta_{\bf ij}|^2 + |\Delta_{\bf ji}|^2
666: \right],\nonumber}
667: \end{array}
668: \end{equation}
669: where
670: $\Delta_{\bf ij}=\langle c_{{\bf j}\downarrow} c_{{\bf j}\uparrow}\rangle$.
671: The effect of $V$ is restricted to generate in-plane superconducting
672: correlations. To further simplify the analysis and to reduce
673: computational efforts, two sublattices are assumed in the $xy$ plane since it is
674: known that this model has a tendency toward antiferromagnetism at half-filling: e.g.,
675: $\langle n_{{\bf i}\uparrow}\rangle=\langle n_{i_z\uparrow}^{(\rm A)}\rangle$
676: and $\langle n_{i_z\uparrow}^{(\rm B)}\rangle$ for $(-1)^{i_x+i_y}=1$
677: (A-sublattice) and $-1$ (B-sublattice), respectively. More details can be
678: found in \cite{alvarez}.
679:
680:
681: To properly
682: describe a stable junction made out of materials with
683: different electronic densities,
684: it is crucial to include long-range Coulomb interactions between electrons
685: and also with
686: the background of positive charges.
687: In this paper, the long-range Coulomb
688: interactions are treated within the Hartree approximation by using the
689: following Poisson's equation~\cite{datta}:
690: \begin{equation}\label{poisson}
691: \nabla^2\phi_{\bf i}
692: = -\alpha\left[\langle n_{\bf i}\rangle-n_+({\bf i})\right],
693: \end{equation}
694: where $\alpha=e/\varepsilon a$ ($\varepsilon$, $e$, and $a$ are the
695: dielectric constant, electronic charge, and lattice spacing, respectively).
696: $n_+({\bf i})$ is the background positive charge density at site {\bf i}.
697: In the following, $e$ and $a$ are set to be 1.
698: Furthermore, $n_+({\bf i})$ is kept constant to a number
699: $n_+^{\rm L}$ ($n_+^{\rm R}$)
700: on the left (right) side of the system. The Poisson's
701: equation (\ref{poisson}) is solved numerically using symmetric discretizations
702: in the $x$ and $y$ directions, and a
703: forward discretization in the $z$ direction, e.g.,
704: $d^2\phi_{\bf i}/dz^2
705: =\phi_{{\bf i}+2{\bf z}}-2\phi_{{\bf i}+{\bf z}}+\phi_{\bf i}$ ({\bf z}: unit
706: vector in $z$ direction). The use of the Hartree approximation here is not
707: a severe problem: the Poisson equation is widely considered to be a reasonable
708: starting point to account for charge transfer, and it is much used in the
709: study of semiconducting systems.
710:
711:
712:
713: To mimic a ferromagnetic half-metallic ground state on the left side of the system,
714: the localized spins $\vec{S}_{\bf i}$ are chosen to be
715: ferromagnetically aligned, pointing
716: into the $z$ direction. To mimic an A-type AF state, these classical spins are
717: chosen with the same orientation in the $xy$ planes but opposite between adjacent planes.
718: On the cuprate side, the mean-field Hamiltonian $H_{\rm MF}$ is first
719: Fourier transformed
720: to momentum space in the $xy$ plane, and the resulting Bogoliubov-De Gennes
721: equation~\cite{bdg} is solved numerically
722: by diagonalizing a $(4L_z\times4L_z)$ Hamiltonian matrix for each momentum $(k_x,k_y)$.
723: The $d$-wave order parameter is
724: $\Delta_d({\bf i})=\delta({\bf i},{\bf x}) + \delta({\bf i},-{\bf x})
725: - \delta({\bf i},{\bf y}) - \delta({\bf i},-{\bf y})$, where
726: $\delta({\bf i},{\bf j}) =
727: \langle c_{{\bf i}\uparrow}c_{{\bf i}+{\bf j}\downarrow} -
728: c_{{\bf i}\downarrow}c_{{\bf i}+{\bf j}\uparrow} \rangle /2$, and
729: {\bf x} and {\bf y} are unit vectors in the
730: $x$ and $y$ directions, respectively.
731:
732:
733: \subsection{FM-SC interface: charge transfer leading to AF}
734: The description of the numerical results starts with a qualitative
735: reproduction of the recent experiments \cite{varela}, where transfer
736: of charge from LCMO to YBCO was found to destroy superconductivity at the interface.
737: In Fig.~\ref{fm_af_sc}, the case when a FM
738: metal forms an interface with a superconducting state (as in Fig.~\ref{systems}(a)) is studied.
739: The work functions and
740: parameters are chosen such that a transfer of charge from the left to the right takes
741: place, as shown in Fig.~\ref{fm_af_sc}(a) where an accumulation of charge at the
742: interface is found. The band-bending picture, which appears to be true
743: not only for doped band insulators but also for Hubbard insulators \cite{nagaosa},
744: suggests that there will be a finite region where the density will be
745: $n$=1, with the chemical potential moving across the Hubbard gap.
746: In Fig.~\ref{fm_af_sc}(b) the local magnetization is shown. On the
747: left, it is FM as expected in a one-orbital manganite model away from $n$=0 or 1.
748: On most of the right, the magnetization is zero as in
749: a regular superconductor. However, in the interface region the magnetization develops a
750: staggered character suggesting the presence of AF order.
751: Thus, as expected from the mean-field formalism the density $n=1$ is associated
752: with antiferromagnetism.
753: Figure~\ref{fm_af_sc}(c) shows the suppression of the superconducting order parameter at the interface.
754: Overall, these simple results reproduce properly the experiments where
755: it was observed that transfer of charge from a manganite LCMO to
756: YBCO led to the suppression of superconductivity at the interface \cite{varela}.
757:
758: \begin{figure}[hbt]
759: \includegraphics[clip=true,width=7.5cm,angle=-0]{FM_SC_2.eps}
760: \begin{center}
761: \caption{
762: (Color online) Transfer of charge from a FM half-metal to a SC state, inducing
763: an AF interface in the process, to reproduce recent experiments \cite{varela}.
764: Shown are the (a) electronic density $n_{\bf i}$,
765: (b) magnetization
766: $m_z({\bf i})=n_{{\bf i}\uparrow}-n_{{\bf i}\downarrow}$, and (c) $d$-wave
767: superconducting order parameter $\Delta_d({\bf i})$, as a function of layer
768: position $i_z$ for two different sublattices [$(-1)^{i_x+i_y}=\pm1$] in the $xy$
769: plane (denoted by circles and crosses). Here ${\bf i}=(i_x,i_y,i_z)$ is a
770: lattice site position. The model parameters used are $J_H=8t$, $t_z=t$,
771: $W_{\rm L}=10t$, and $n_+^{\rm L}=0.7$ for the left side of the system, and
772: $U=4t$, $V=-3t$, $t_z=0.1t$ (to simulate weak hopping between the Cu oxide
773: layers), $W_{\rm R}=0$, and $n_+^{\rm R}=0.7$ for the
774: right side of the system. $\alpha=1$ is set for the whole system which has a size
775: $L=16\times16\times24$. The interface is located at $i_z=12.5$.
776: The localized spins in the left side of the system are fixed to be
777: ferromagnetic. The temperature considered here is very low, $T$=$t$/400.
778: }
779: \label{fm_af_sc}
780: \end{center}
781: \end{figure}
782:
783: \subsection{Manganite-Cuprate interface, leading to electron-doped SC}
784:
785: After having crudely reproduced the main qualitative aspects of previous
786: experiments, the same formalism can be used to
787: formulate predictions for other systems. Of particular interest is the
788: combination of an A-type AF state (as it occurs for instance in undoped or highly doped LSMO
789: and also in bilayer compounds \cite{dagotto-CMR,book})
790: and an $undoped$ cuprate.
791: If the transfer of charge occurs in the same directions as before, as
792: already suggested by the discussion on experimental work functions,
793: then it would be expected
794: that at the interface a density larger than $n$=1 would be produced (see the sketch in Fig.~\ref{systems}(b)).
795: If this doping is as large as 5-10\%, then an electron doped
796: superconductor could be created in a real system.
797:
798: The actual calculations are conceptually simple and the results are shown in
799: Fig.~\ref{a-af_sc}.
800: In (a), the density is presented: once again
801: a robust region in parameter space is identified where the transfer of
802: charge occurs from the AF manganite to the AF cuprate, leading in this case to $n$ larger
803: than 1 at the interface. In (b), the local magnetization is shown.
804: In this case, the G-type AF order is seen on most of the right (cuprate) region, but
805: this magnetic order disappears at the interface due to the transfer of charge. Finally, in (c)
806: the superconducting order parameter is shown to become
807: nonzero at the interface. Thus,
808: as anticipated from the introduction, an electron-doped superconductor is predicted to occur
809: in the manganite/cuprate system described here, within a simple mean-field approximation, for a pair of oxides
810: with the proper location of chemical potentials.
811:
812: If the A-type AF manganite state for the left side is replaced by a FM state, as before,
813: then a similar transfer of charge occurs and within mean-field a SC state is also generated.
814: However, it is known that the proximity-effect influence of ferromagnetism is detrimental
815: to singlet superconductivity, thus for experimental realizations of this scenario, a
816: non-FM state appears more suitable.
817:
818:
819: {\it Note}: if the transfer of charge would occur in the other direction, namely
820: from the $n$=1 AF to a doped FM material, then the cuprate interface would have less
821: charge and a narrow layer with the properties of a $hole-doped$ cuprate
822: could be expected. Thus, the simple motive of this effort
823: works both ways at the mean-field level.
824: However, the analysis of work functions from the experimental viewpoint, already discussed,
825: suggests that in order to generate interfacial superconductivity involving a manganite,
826: electron doping is the most likely outcome since the chemical potential of LSMO is
827: above that of LCO, NCO, SCO, and YBCO.
828:
829:
830:
831:
832:
833:
834: \begin{figure}[hbt]
835: \includegraphics[clip=true,width=7.5cm,angle=-0]{A-AF_SC.eps}
836: \begin{center}
837: \caption{
838: (Color online) Transfer of charge from an A-type AF state (as it
839: occurs in some doped manganites \cite{dagotto-CMR})
840: to an AF insulator (LCO, SCO, NCO, or YBCO), inducing
841: an electron-doped SC state at the interface.
842: The actual
843: model parameters
844: used here are $J_H=8t$, $t_z=t$, $W_{\rm L}=14t$, and $n_+^{\rm L}=0.7$ for
845: the left side of the system, and $U=4t$, $V=-3t$, $t_z=0.1t$,
846: $W_{\rm R}=0$, and $n_+^{\rm R}=1.0$ for the right side of the system.
847: $\alpha=1$ is set for the whole system, with
848: $L=16\times16\times24$ being the lattice studied.
849: The interface is located at $i_z=12.5$.
850: The localized spins in the left side of the systems are fixed to be
851: antiferromagnetic in an A-type state, and the temperature of the study was $T$=$t$/400.
852: }
853: \label{a-af_sc}
854: \end{center}
855: \end{figure}
856:
857:
858:
859: \subsection{SC at BI-AF and Metal-AF interfaces}
860:
861: \begin{figure}[hbt]
862: \includegraphics[clip=true,width=7.5cm,angle=-0]{FM_SC_3.eps}
863: \begin{center}
864: \caption{
865: (Color online) Generation of a SC state at the interface between a band
866: insulator (BI) and an AF insulator and also between a standard metal and an AF.
867: As in other figures, shown are the
868: electron density $n_{\bf i}$
869: [(a) and (c)], and staggered
870: magnetization $(-1)^{i_z}m_z({\bf i})$ and $d$-wave superconducting order
871: parameter $\Delta_d({\bf i})$ [(b) and (d)] as a function of layer position
872: $i_z$.
873: The model parameters used in (a) and (b) are $J_H=0$ (to avoid ferromagnetism), $t_z=t$,
874: $W_{\rm L}=0$, and $n_+^{\rm L}=0.0$ (to get a crude band insulator)
875: for the left side of the system, and
876: $U=4t$, $V=-3t$, $t_z=0.1t$, $W_{\rm R}=0$, and $n_+^{\rm R}=1.0$
877: for the right
878: side of the system. The same parameters are used in (c) and (d) except
879: $n_+^{\rm L}=0.4$ for the left side of the system.
880: $\alpha=1$ is set for the whole system with
881: $L=16\times16\times24$ being the lattice studied,
882: and the interface is located at $i_z=12.5$.
883: }
884: \label{bi_af}
885: \end{center}
886: \end{figure}
887:
888: The generation of a superconductor via transfer of charge from another
889: compound is certainly not restricted to occur only when LSMO is involved, but
890: it should happen under far more general circumstances. For instance,
891: the case of a band-insulator (BI) forming an interface with an AF
892: was also studied.
893: Figure~\ref{bi_af}(a) shows the density profile, indicating that the parameters of
894: the calculation are such that the
895: transfer of charge this time occurs from the AF to the BI, inducing
896: a region in the cuprate with hole doping, and concomitant superconductivity
897: (as shown in (b)). Certainly, electron-doped SC can be induced as well by
898: adjusting the work functions.
899:
900: If instead of a BI, a standard metal is used
901: (modeled in our study by merely using Hubbard $U$=0 in a tight-binding model),
902: then for appropriate work functions also a charge transfer is to be expected
903: (see (c) and (d)). Thus, the most important aspect of the problem is
904: to identify materials with the proper relative location of work functions such that the
905: charge transfer occurs in the proper direction, and also such that there
906: is a good matching between lattice spacings to avoid generating extra
907: complications for the transfer of charge to occur.
908:
909: \section{Extended Dynamical Mean Field Theory Results for Hubbard Model Interfaces}
910:
911:
912: To place our findings on a firmer footing, we describe here
913: the results obtained from cluster dynamical mean field theory
914: (CDMFT) \cite{kotliar2001} generalized to a layered geometry for the
915: case of a ferromagnetic -- Mott insulator
916: interface. The previous section showed that there are many similarities
917: in the use of FM or AF manganites in the heterostructure, since their main role is the donation of
918: carriers to the cuprate. Since technically the case of FM is simpler, in this
919: section we focus on a FM manganite.
920: CDMFT is a powerful technique which maps a full many-body
921: problem onto that of a cluster embedded in a self-consistent
922: medium. The model Hamiltonian considered here is identical to
923: Eqs.~1 and 2, with the addition of the local Coulomb interaction represented by the
924: Hubbard $U$ on the Mott insulating side of the interface (and without the
925: explicitly attractive nearest-neighbors attraction) . Long-range
926: Coulomb interactions are treated again using the Poisson equation as
927: in Eq.~4.
928:
929: CDMFT can be generalized to treat the effects of a layered geometry in
930: the same manner as DMFT was adapted to treat a film
931: geometry \cite{potthoff2003,okamoto2004}. Each layer is mapped onto an
932: independent cluster impurity embedded in a medium. The solution for
933: the self-energy of each layer is then used via a self-consistency
934: condition to obtain the local Green's function of the entire
935: lattice. Hopping between the layers is taken into account only at the
936: level of the self-consistency condition. To make the problem
937: computationally feasible, we fix the number of layers to $L_{\rm
938: FM}=5$ and $L_{\rm MI}=5$ for the ferromagnet and the Mott insulator,
939: respectively. Further, each layer is mapped to a 4-site cluster
940: embedded in a bath of 8 sites. SC correlations are included in the
941: bath around the MI layers to allow for the SC instability and a paramagnetic
942: solution is imposed. Individual
943: layers are oriented in the $xy$-plane and stacked in the
944: $z$-direction. The algorithm starts by making an initial guess for the
945: local chemical potential and the bath parameters across the layered
946: structure. Solving the cluster impurity model in each layer using the
947: Lanczos method we obtain the cluster Green's function, cluster density,
948: and self-energy. The density in each layer is used to obtain the new
949: local potential $\phi_z$ using the Poisson equation (Eq.~4). The
950: self-energy in each layer is used to obtain a local Green's function
951: for the lattice using the equation below:
952: \begin{equation}
953: G_{\rm loc}(z,\omega)=\int\frac{dK_xdK_y}{\pi^2}G(z,z',K_x,K_y,\omega)
954: \end{equation}
955: where
956: \begin{equation}
957: \begin{array}{l}
958: \displaystyle{
959: G(z,z',K_x,K_y,\omega)=
960: } \\
961: \displaystyle{
962: [\omega+\mu_z-t(K_x,K_y,z,z')-\Sigma(z,z',K_x,K_y,\omega)]^{-1}.
963: }
964: \end{array}
965: \end{equation}
966: Here, $t(K_x,K_y,z,z')$ denotes the Fourier transform of the hopping
967: matrix (both intralayer and interlayer), $K_x$ and $K_y$ denote the
968: superlattice momenta within each layer, and $z$ denotes the layer
969: index. $\mu_z$ includes all the terms that couple to the density,
970: including the chemical potential, $\phi_z$, site potential,
971: and the work function on either side of the interface.
972: The Green's function $G_{\rm loc}(z,\omega)$ is
973: combined with the Dyson equation to give a new Weiss field that is
974: then used to obtain a new set of bath parameters using a conjugate
975: gradient minimization. Convergence is achieved when the bath
976: parameters for each layer and the density profile across the layers do
977: not change with further iterations.
978: \begin{figure}[ht]
979: \begin{center}
980: \includegraphics[width=7.0cm,angle=-0] {FM-MI-CDMFT.eps}
981: \end{center}
982: \caption{Transfer of charge from a FM manganite to a Mott insulating
983: cuprate inducing a $d$-wave SC
984: state at the interface in a 10-layer heterostructure, obtained using CDMFT
985: generalized to a layered geometry. (a) is the electronic density profile
986: $n_z$; (b) is the magnetic moment in each layer $z$, denote by $m_z(i_z)$;
987: and (c) is the $d$-wave
988: SC order parameter, $\Delta_d(i_z)=\langle c_{z,i\uparrow}c_{z,i+x\downarrow}\rangle$
989: across the heterostructure. The model parameters used are $J_H=8t$, $t_z=t$, $W_L=12t$,
990: $n_{+}^{L}=0.7$, $n_{+}^{R}=1.0$, $W_R=0$, $\alpha=1$, and $U/t$=8.}
991: \label{interface}
992: \end{figure}
993:
994: In Fig.~\ref{interface}(a) the density profile is presented. The bulk
995: density is at $n$=0.7 on the left side (Mn oxide)
996: and $n$=1.0 on the right (Cu oxide, $U/t$=8). As expected from
997: the previous discussions, electron doping of the Mott insulator
998: occurs at the interface due to a transfer of charge
999: from the manganite to the cuprate.
1000: Figure~\ref{interface}(b) provides the magnetization at each layer. Note
1001: that the effects of penetration of the FM moment onto the Mott insulator side are not considered
1002: in this calculation. In a real setup, it is likely that a finite polarization will be found
1003: on the right hand side of the figure.
1004: Figure~\ref{interface}(c)
1005: shows the $d$-wave SC order parameter, which in this heterostructure
1006: becomes non-zero in the cuprate's
1007: interfacial region and decays rapidly into the bulk of the Mott insulator. Thus, an
1008: electron-doped $d$-wave superconductor can be realized at the interface
1009: of a manganite and a cuprate, due to the expected transfer of charge
1010: between them, even when dynamical effects in the mean-field approximation
1011: are taken into account.
1012:
1013:
1014:
1015:
1016:
1017:
1018:
1019:
1020: \section{Discussion and Summary}
1021:
1022: In this manuscript, the possible charge transfer from a manganite to an undoped cuprate was
1023: discussed in the context of oxide heterostructures. This issue is nontrivial since $a$ $priori$
1024: the existence of a robust gap in the undoped
1025: Cu oxides would have suggested the lack of available states for manganite electrons to pour
1026: into the antiferromagnetic cuprates near the interfaces. The recently discovered $indirect$ nature
1027: of the Cu-oxides gaps effectively reduces the magnitude of the Hubbard gap. Taking this into
1028: consideration, our analysis suggests that the Fermi level of manganites, in a robust range of hole
1029: doping, lies above the chemical potential of several cuprates and even above the bottom of the upper
1030: Hubbard band of SCO and NCO.
1031: This opens the possibility
1032: of electron doping of high-$T_c$'s at interfaces with manganites (and other oxides).
1033: Under idealized conditions, the doping discussed here may induce a superconducting state (if the
1034: manganite is not ferromagnetic), an exotic metallic state containing polarized carriers in an
1035: antiferromagnetic background (if the manganite is ferromagnetic),
1036: or still a superconducting state,
1037: likely with a triplet component, if the spin polarization of the carriers coming from the manganite
1038: is only partial.
1039:
1040: An important component of our effort is the introduction of a systematic procedure to analyze
1041: photoemission and diffusion voltage experiments to predict the direction of flow of charge at
1042: interfaces. A figure with the relative Fermi levels and gap locations of several oxides was presented.
1043:
1044: If the superconducting state is ever realized in clean interfaces, a fundamental issue to address
1045: is the unveiling of the phase diagram of cuprates in the absence of quenched disorder. This is true for
1046: both electron and hole doping. Recent phenomenological calculations \cite{alvarez} involving noninteracting carriers
1047: in interaction with the AF and SC order parameters (an extension of the traditional Landau-Ginzburg approach)
1048: suggest that these phases should be separated by either a region of local coexistence or a first-order
1049: transition. A glassy state, as in the widely studied phase diagram of LSCO, was reproduced $only$ incorporating
1050: quenched disorder. For a sketch of these diagrams see Fig.~\ref{phase_SC}. The generation of superconductivity
1051: at interfaces may reveal the true phase diagram of clean cuprates.
1052:
1053: \begin{figure}[hbt]
1054: \includegraphics[clip=true,width=7.5cm,angle=-0]{phase_SC.eps}
1055: \begin{center}
1056: \caption{ (a) and (b) are the possible phase diagrams of the cuprates in the clean limit \cite{alvarez},
1057: which may be experimentally realized at the interfaces discussed in this paper. No distinction is
1058: made between hole or electron doping, $x$ represents both. (c) is the well-known phase diagram of chemically
1059: doped LSCO. According to Ref.~\cite{alvarez}, the glassy state between AF and SC phases is caused by
1060: quenched disorder.
1061: }
1062: \label{phase_SC}
1063: \end{center}
1064: \end{figure}
1065:
1066: It is clear that our calculations can only be considered as suggestive, and its main purpose is to induce further
1067: work in this area. Important simplifications employed in the study,
1068: including the neglect of lattice reconstructions, vacancies, and polarity effects, need to be addressed.
1069: $Ab$ $initio$ calculations are crucial to clarify these issues. And, of course, the experimental
1070: realization of the interfaces discussed here would provide a definitive answer to the proposal
1071: of electron doping of cuprates at the interfaces.
1072:
1073:
1074: \section{Acknowledgment}
1075: We thank I. Bozovic, H. Y. Hwang, M. Kawasaki, T. Kopp,
1076: H. Kumigashira, S. Pennycook, A. Sawa, Y. Tokura,
1077: and M. Varela for valuable discussions. This work was supported
1078: in part by the NSF grant DMR-0443144 and by the Division of Materials Sciences and
1079: Engineering, Office of Basic Energy Sciences, U.S. Department of Energy,
1080: under contract DE-AC05-00OR22725 with Oak Ridge National Laboratory,
1081: managed and operated by UT-Battelle, LLC.
1082: Support by the LDRD program at ORNL is also acknowledged.
1083:
1084:
1085: %%%%%%%%%%%%%%%%%
1086: %% Reference
1087: %%%%%%%%%%%%%%%%
1088:
1089: \begin{thebibliography}{99}
1090: \bibitem{ohtomo02} A. Ohtomo, D. A. Muller, J. L. Grazul, and H. Y. Hwang, Nature {\bf 419}, 378 (2002).
1091:
1092: \bibitem{ohtomo04}A. Ohtomo and H. Y. Hwang, Nature {\bf 427}, 423 (2004).
1093:
1094: \bibitem{okamoto04a} S. Okamoto and A. Millis, Nature {\bf 428}, 630 (2004).
1095:
1096: \bibitem{okamoto04b} S. Okamoto and A. J. Millis, Phys. Rev. B {\bf 70}, 075101 (2004).
1097:
1098: \bibitem{okamoto2004} S. Okamoto and A. J. Millis, Phys. Rev. B
1099: {\bf 70}, 241104(R) (2004).
1100:
1101: \bibitem{nakagawa}N. Nakagawa, H. Y. Hwang, and D. A. Muller, Nature Materials {\bf 5}, 204 (2006),
1102: and references therein.
1103:
1104: \bibitem{thiel} S. Thiel, G. Hammerl, A. Schmehl, C. W. Schneider, and J. Mannhart, Science {\bf 313}, 1942 (2006),
1105: and references therein.
1106:
1107: \bibitem{dagotto} E. Dagotto, Science {\bf 309}, 257 (2005), and references therein.
1108:
1109:
1110: \bibitem{yoshida} A. F. Volkov, F. S. Bergeret and K. B. Efetov,
1111: Phys. Rev. Lett. {\bf 90}, 117006 (2003). See also
1112: N. Yoshida, M. Fogelstrom, cond-mat/0511009.
1113:
1114: \bibitem{eschrig} M. Eschrig, J. Kopu, J. C. Cuevas, and Gerd Schon, Phys. Rev. Lett. {\bf 90}, 137003 (2003).
1115:
1116: \bibitem{kontos} T. Kontos, M. Aprili, J. Lesueur, F. Genet, B. Stephanidis, and R. Boursier, Phys. Rev. Lett. {\bf 89}, 137007 (2002).
1117:
1118: \bibitem{vasko}
1119: V. A. Vas'ko, V. A. Larkin, P. A. Kraus, K. R. Nikolaev, D. E. Grupp, C. A. Nordman, and A. M. Goldman,
1120: Phys. Rev. Lett. 78, 1134 (1997).
1121:
1122: \bibitem{melo} C. A. R. Sa de Melo, Phys. Rev. Lett. {\bf 79}, 1933 (1997).
1123:
1124: \bibitem{dagotto-CMR} E. Dagotto, T. Hotta, and A. Moreo, Physics Reports {\bf 344}, 1 (2001).
1125:
1126: \bibitem{book} E. Dagotto, {\it Nanoscale Phase Separation and Colossal Magnetoresistance}, Springer (2002).
1127:
1128:
1129: \bibitem{mannhart} J. Mannhart, D. G. Schlom, J. G. Bednorz, and
1130: K. A. Muller, Phys. Rev. Lett. {\bf 67}, 2099 (1991).
1131:
1132: \bibitem{ahn} C. H. Ahn, S. Gariglio, P. Paruch, T. Tybell,
1133: L. Antognazza, and J.-M. Triscone, Science {\bf 284}, 1152 (1999).
1134:
1135: \bibitem{pavlenko} For recent work in this area see
1136: N. Pavlenko, I. Elfimov, T. Kopp, and G. A. Sawatzky, cond-mat/0605589, and references therein.
1137: For related literature see: V. Koerting, Q. Yuan, P.J. Hirschfeld, T. Kopp, and J. Mannhart,
1138: Phys. Rev. B {\bf 71}, 104510 (2005); N. Pavlenko and T. Kopp,
1139: Phys. Rev. B {\bf 72}, 174516 (2005).
1140:
1141:
1142: \bibitem{varela} M. Varela, A.R. Lupini, V. Pe\~na, Z. Sefrioui, I. Arslan,
1143: N.D. Browning, J. Santamaria, S.J. Pennycook, cond-mat/0508564; see also
1144: Todd Holden, H.-U. Habermeier, G. Cristiani, A. Golnik, A. Boris, A. Pimenov,
1145: J. Humlicek, O. I. Lebedev, G. Van Tendeloo, B. Keimer, and C. Bernhard,
1146: Phys. Rev. B {\bf 69}, 064505 (2004);
1147: A. Hoffmann, S. G. E. te Velthuis, Z. Sefrioui, J. Santamaria,
1148: M. R. Fitzsimmons, S. Park, and M. Varela,
1149: Phys. Rev. B{\bf 72}, 140407(R) (2005);
1150: V. Pe\~na, C. Visani, J. Garcia-Barriocanal, D. Arias, Z. Sefrioui,
1151: C. Leon, J. Santamaria, and C. A. Almasan,
1152: Phys. Rev. B {\bf 73}, 104513 (2006).
1153:
1154: \bibitem{hilgenkamp} H. Hilgenkamp and J. Mannhart, Rev. Mod. Phys. {\bf 74}, 485 (2002); and references therein.
1155: See also B. Nikolic, J. K. Freericks, and P. Miller, Phys. Rev. B {\bf 65}, 064529 (2002); U. Schwingenschlogl and
1156: C. Schuster, cond-mat/0702098.
1157:
1158: \bibitem{sefrioui} Z. Sefrioui, D. Arias, V. Pe\~na, J. E. Villegas, M. Varela, P. Prieto, C. Le\'on, J. L. Martinez,
1159: and J. Santamaria, Phys. Rev. B {\bf 67}, 214511 (2003). See also V. Pe\~na, Z. Sefrioui, D. Arias, C. Le\'on, J. Santamaria,
1160: J. L. Martinez, S. G. E. te Velthuis, and A. Hoffmann, Phys. Rev. Lett. {\bf 94}, 057002 (2005); J. Chakhalian, J. W. Freeland,
1161: G. Srajer, J. Strempfer, G. Khaliullin, J. C. Cezar, T. Charlton, R. Dalgliesh, C. Bernhard, G. Cristiani, H-U. Habermeier, and
1162: B. Keimer, Nature Physics {\bf 2}, 244 (2006).
1163:
1164: \bibitem{oh} The $hole$ doping of an undoped cuprate when in combination with STO has already been reported in
1165: S. Oh, M. Warusawithana, and J. N. Eckstein, Phys. Rev. B {\bf 70}, 064509 (2004). See also
1166: J. N. Eckstein and I. Bozovic, Annu. Rev. Mater. Sci. {\bf 25}, 679 (1995); G. Yu. Logvenov, A. Sawa, C. W. Schneider and
1167: J. Mannhart, App. Phys. lett. {\bf 83}, 3528 (2003); and references therein.
1168:
1169: \bibitem{brey} L. Brey, cond-mat/0611594, and references therein.
1170:
1171:
1172: \bibitem{Schulte01}K. Schulte, M. A. James, L. H. Tjeng, P. G. Steeneken, G. A. Sawatzky, R. Suryanarayanan, G. Dhalenne, and A. Revcolevschi,
1173: Phys. Rev. B {\bf 64}, 134428 (2001).
1174:
1175: \bibitem{Suzuki89}M. Suzuki, Phys. Rev. B {\bf 39}, 2312 (1989).
1176:
1177: \bibitem{Ino00} A. Ino, C. Kim, M. Nakamura, T. Yoshida, T. Mizokawa, Z-X. Shen, A. Fujimori, T. Kakeshita, H. Eisaki,
1178: and S. Uchida, Phys. Rev. B {\bf 62}, 4137 (2000).
1179:
1180: \bibitem{polarons}O. R{\"o}sch, O. Gunnarsson, X. J. Zhou, T. Yoshida, T. Sasagawa, A. Fujimori,
1181: Z. Hussain, Z.-X. Shen, and S. Uchida, Phys. Rev. Lett. {\bf 95}, 227002 (2005).
1182:
1183: \bibitem{Allen90}J. W. Allen, C. G. Olson, M. B. Maple, J.-S. Kang, L. Z. Liu, J.-H. Park, R. O. Anderson, W. P. Ellis, J. T. Markert,
1184: Y. Dalichaouch, and R. Liu, Phys. Rev. Lett. {\bf 64}, 595 (1990).
1185:
1186: \bibitem{Namatame90}H. Namatame, A. Fujimori, Y. Tokura, M. Nakamura, K. Yamaguchi, A. Misu, H. Matsubara, S. Suga, H. Eisaki,
1187: T. Ito, H. Takagi, and S. Uchida, Phys. Rev. B {\bf 41}, 7205 (1990).
1188:
1189: \bibitem{alternative} Note that the chemical potential difference between
1190: LCO and NCO is still controversial. Other estimations
1191: arrive to a value $\sim 1$ eV
1192: for this quantity. See P. G. Steeneken, L. H. Tjeng, G. A. Sawatzky, A. Tanaka,
1193: O. Tjernberg, G. Ghiringhelli, N. B. Brookes, A. A. Nugroho, and A. A. Menovsky,
1194: Phys. Rev. Lett. {\bf 90}, 247005 (2003).
1195:
1196: \bibitem{Fujimori02}A. Fujimori, A. Ino, J. Matsuno, T. Yoshida, K. Tanaka, and T. Mizokawa,
1197: J. Electr. Spectrosc. Relat. Phenom. {\bf 124}, 127 (2002).
1198:
1199: \bibitem{Tajima89}S. Tajima, H. Ishii, T. Nakamura,H. Takagi, S. Uchida, M. Seki, S. Suga, Y. Hidaka, M. Suzuki, T. Murakami,
1200: K. Oka, and H. Unoki, J. Opt. Soc. Am. B {\bf 6}, 475 (1989).
1201:
1202: \bibitem{Tsutsui99}K. Tsutsui, T. Tohyama, and S. Maekawa, Phys. Rev. Lett. {\bf 83}, 3705 (1999).
1203:
1204: \bibitem{Hasan00}M. Z. Hasan, E. D. Isaacs, Z.-X. Shen, L. L. Miller, K. Tsutsui, T. Tohyama, and S. Maekawa, Science {\bf 288}, 1811 (2000).
1205:
1206: \bibitem{BIS98} A. Fujimori, A. Ino, T. Mizokawa, C. Kim, Z.-X. Shen,
1207: T. Sasagawa, T. Kimura, K. Kishio, M. Takaba, K. Tamasaku, H. Eisaki
1208: and S. Uchida, J. Phys. Chem. Solids {\bf 59}, 1892 (1998).
1209:
1210: \bibitem{Muraoka04}Y. Muraoka, T. Muramatsu, J. Yamaura, and Z. Hiroi, Appl. Phys. Lett. {\bf 85}, 2950 (2004).
1211:
1212: \bibitem{Nakamura07}M. Nakamura, A. Sawa, H. Sato, H. Akoh, M. Kawasaki, and Y. Tokura, Phys. Rev. B {\bf 75}, 155103 (2007).
1213:
1214: \bibitem{SawaPC}T. Fujii, M. Kawasaki, A. Sawa, Y. Kawazoe, H. Akoh, and Y. Tokura, Phys. Rev. B (2007), in press.
1215:
1216: \bibitem{Muramatsu05}T. Muramatsu, Y. Muraoka, and Z. Hiroi, Jpn. J. Appl. Phys. {\bf 44}, 7367 (2005).
1217:
1218: \bibitem{Matsuno02}J. Matsuno, A. Fujimori, Y. Takeda and M. Takano, Europhys. Lett. {\bf 59}, 252 (2002).
1219:
1220: \bibitem{Ino97}A. Ino, T. Mizokawa, A. Fujimori, K. Tamasaku, H. Eisaki, S. Uchida, T. Kimura, T. Sasagawa, and K. Kishio,
1221: Phys. Rev. Lett. {\bf 79}, 2101 (1997).
1222:
1223: \bibitem{Harima03}N. Harima, A. Fujimori, T. Sugaya, and I. Terasaki, Phys. Rev. B {\bf 67}, 172501 (2003).
1224:
1225: \bibitem{Yagi06}H. Yagi, T. Yoshida, A. Fujimori, Y. Kohsaka, M. Misawa, T. Sasagawa, H. Takagi, M. Azuma and M. Takano, Phys. Rev. B {\bf 73}, 172503 (2006).
1226:
1227: \bibitem{Harima01}N. Harima, J. Matsuno, A. Fujimori, Y. Onose, Y. Taguchi, and Y. Tokura, Phys. Rev. B {\bf 64}, 220507(R) (2001).
1228:
1229: \bibitem{Jong03}M. P. de Jong, V. A. Dediu, C. Taliani, and W. R. Salaneck, J. Appl. Phys. {\bf 94}, 7292 (2003).
1230:
1231: \bibitem{Minohara07}M. Minohara, I. Ohkuho, H. Kumigashira, and M. Oshima,
1232: Appl. Phys. Lett. {\bf 90}, 132123 (2007).
1233:
1234: \bibitem{Reagor04}W. Reagor, S. Y. Lee, Y. Li, and Q. X. Jia, J. Appl. Phys. {\bf 95}, 7971 (2004).
1235:
1236: \bibitem{YBCO-work} The work function of YBCO has been experimentally measured. According to T. Hirano,
1237: M. Ueda, K. Matsui, T. Fujii, K. Sakuta, and T. Kobayashi, Jpn. J. Appl. Phys. {\bf 31}, L1345 (1992),
1238: the YBCO work function is 6.1 eV. Since the work function of LSMO
1239: is $\sim 4.8$~eV \cite{Jong03,Minohara07,Reagor04}, their difference is 1.3 eV, which is not too different (given
1240: the uncertainties in these experimental results) from the
1241: numbers deduced from Fig.~\ref{fig:LCO_SCO_NbSTO_YBCO_LSMO}.
1242:
1243:
1244: \bibitem{STO-comment}
1245: Are there other transition-metal oxides that can replace manganites in this context?
1246: Consider first the case of ruthenates. In addition to the $3d$ oxides, clearly the
1247: $4d$ transition-metal oxides are another potentially important class of materials for our purposes.
1248: Among them, SrRuO$_3$ (SRO) is known to be highly conductive and, therefore, a potential candidate
1249: for electrode in transition-metal based devices.
1250: However,
1251: through permittivity measurements, the work function of (001) SRO is estimated to be 5.2~eV (see
1252: X. Fang and T. Kobayashi, Appl. Phys. A {\bf 69}, S587 (1999)).
1253: This SRO work function is rather large compared with the value
1254: $\sim 3.5$~eV of the double-layered La$_{1.2}$Sr$_{1.8}$Mn$_2$O$_7$
1255: and $\sim4.8$ of the cubic perovskite LSMO \cite{Minohara07}.
1256: This may indicate the following trend: the larger the atomic number of the transition-metal is,
1257: the larger the work function becomes.
1258: In other words, if one wants to inject electrons to some material, one may need to
1259: use transition-metal oxides with $smaller$ atomic number than manganites (since reducing the
1260: atomic number also reduces the electro-negativity)
1261: From this point of view, electron-rich titanates (via the addition of Nb)
1262: can provide another route for electron doping of cuprate parent compounds,
1263: as very recently experimentally reported \cite{Nakamura07}.
1264: Our study is by no means restricted to manganites as electron
1265: donors.
1266:
1267:
1268: \bibitem{tuv} E. Dagotto, J. Riera, Y. C. Chen, A. Moreo, A. Nazarenko,
1269: F. Alcaraz, and F. Ortolani, \prb {\bf 49}, 3548 (1994); and
1270: references therein. See also R. Micnas, J. Ranninger, and S. Robaszkiewicz, Rev. Mod. Phys. {\bf 62},
1271: 113 (1990); I. Martin, G. Ortiz, A. V. Balatsky, and A. R. Bishop, Int. J. of Mod. Phys. {\bf 14}, 3567 (2000).
1272:
1273: \bibitem{alvarez} G. Alvarez, M. Mayr, A. Moreo, and E. Dagotto, Phys. Rev. B {\bf 71}, 014514 (2005). See also
1274: M. Mayr, G. Alvarez, A. Moreo, and E. Dagotto, Phys. Rev. B {\bf 73}, 014509 (2006).
1275:
1276:
1277: \bibitem{datta} See, for example, S. Datta,
1278: {\it Electronic transport in mesoscopic systems}
1279: (Cambridge University Press, Cambridge, 1997).
1280:
1281: \bibitem{bdg} See, e.g., P. G. De Gennes
1282: {\it Superconductivity of Metals and Alloys}
1283: (Perseus Books, New York, 1999).
1284:
1285:
1286: \bibitem{nagaosa} T. Oka and N. Nagaosa,
1287: Phys. Rev. Lett. {\bf 95}, 266403 (2005).
1288:
1289:
1290: \bibitem{kotliar2001} G. Kotliar, S. Y. Savrasov, G. Palsson, and G. Biroli
1291: Phys. Rev. Lett. {\bf 87}, 186401 (2001).
1292:
1293: \bibitem{potthoff2003} M. Potthoff and W. Nolting, Phys. Rev. B {\bf 59}, 2549;
1294: {\bf 60}, 7834 (1999);
1295: S. Schwieger, M. Potthoff, and W. Nolting,
1296: Phys. Rev. B {\bf 67}, 165408 (2003).
1297:
1298:
1299:
1300:
1301:
1302:
1303:
1304:
1305:
1306:
1307:
1308:
1309:
1310:
1311:
1312:
1313:
1314:
1315:
1316:
1317:
1318:
1319:
1320:
1321:
1322:
1323:
1324:
1325:
1326:
1327:
1328:
1329: \end{thebibliography}
1330:
1331: \end{document}
1332:
1333: