cond-mat0306569/wat.tex
1: 
2: 
3: \documentstyle[prl,aps,epsfig,multicol]{revtex}
4: %\documentclass[aps,prl,twocolumn]{revtex4}
5: %\textwidth=16truecm
6: %\textheight= 20truecm
7: %\hoffset=-1truecm
8: %\voffset=-2truecm
9: 
10: \begin{document}
11: %\preprint{IMSc-2003/06/16}
12: \def\be{\begin{equation}}
13: \def\ee{\end{equation}}
14: \def\bearr{\begin{eqnarray}}
15: \def\eearr{\end{eqnarray}}
16: \def\tc{$T_c~$}
17: \def\tcl{$T_c^{1*}~$}
18: \def\c2{ CuO$_2~$}
19: \def\ruo{ RuO$_2~$}
20: \def\lsco{LSCO~}
21: \def\bi{bI-2201~}
22: \def\tl{Tl-2201~}
23: \def\hg{Hg-1201~}
24: \def\sro{$Sr_2 Ru O_4$~}
25: \def\rc{$RuSr_2Gd Cu_2 O_8$~}
26: \def\mgb{$MgB_2$~}
27: \def\pz{$p_z$~}
28: \def\ppi{$p\pi$~}
29: \def\sqo{$S(q,\omega)$~}
30: \def\tperp{$t_{\perp}$~}
31: \def\cob{$CoO_2$~}
32: \def\nxcob{$Na_x CoO_2.yH_2O$~}
33: \def\ncob{$Na_{0.5} CoO_2$~}
34: \def\half{$\frac{1}{2}$~}
35: \def\nycob{$A_xCoO_{2+\delta}$~}
36: \def\naxcob{$Na_xCoO_2$~}
37: \def\wat{$H_2O$~}
38: \def\na{$Na$~}
39: \def\ewat{$\epsilon_{\rm H_2O}$~}
40: 
41: 
42: \title{How Ice enables Superconductivity in \nxcob by melting charge order:\\
43: Possibility of novel Electric Field Effects}
44: 
45: \author{ G. Baskaran\\
46: The Institute of Mathematical Sciences\\
47: C.I.T. Campus, Chennai 600 113, India }
48: 
49: %\date{today}
50: 
51: \maketitle
52: 
53: \begin{abstract}
54: 
55: Charge ordering in doped \cob planes near the commensurate fillings
56: $x = \frac{1}{4}$ and $\frac{1}{3}$ are considered for \nxcob and 
57: suggested to be competitors to superconductivity, leading to 
58: the experimentally seen narrow superconducting dome bounded by 
59: commensurate doping: $\frac{1}{4}<x< \frac{1}{3}$. 
60: Intercalated hydrogen bonded \wat 
61: network, by its enhanced dielectric constant, screen and frustrate 
62: local {\em charge order condensation energy} and replace a generic 
63: `charge glass order' by superconductivity in the dome.  An access 
64: to superconductivity and charge order, available through the new 
65: water channel, is used to predict novel effects such as 
66: `Electrical Modulation of Superconductivity' and `Electroresistance
67: Effect'.
68: \end{abstract}
69: 
70: %\pacs{PACS number: 13.20.He, 11.30.Er, 12.15Hh, 11.80Et}
71: 
72: %\maketitle
73: 
74: \begin{multicols}{2}[]
75: 
76: Discovery of superconductivity in \nxcob by Takada and 
77: collaborators\cite{takada} have opened the possibility of realizing
78: unconventional superconductivity and novel quantum states in 2D arising 
79: from strong electron correlations in doped \cob layers. Water of a 
80: right proportion ($y \approx \frac{4}{3}$) seems absolutely 
81: necessary\cite{cava1,cava2,chu,jin} for stabilizing 
82: superconductivity, suggesting $H_2O$'s critical role. While water does 
83: wonders in nature, its key role here is some what puzzling. Elucidating 
84: its role in this unusual superconductor is an important 
85: task from material science and physics point of view.  This is what 
86: the present paper attempts using phenomenological and theoretical
87: considerations.
88: 
89: Enthused by the remarkable discovery of superconductivity in \nxcob,
90: the present author\cite{gbcob} and others\cite{rvbcob}
91: have suggested a single band t-J model
92: as an appropriate model to understand superconductivity and low energy 
93: electronic phenomena. A phase diagram has been suggested using ideas 
94: of resonating valence bond (RVB) theory developed for 
95: cuprates\cite{pwagbrvb}. A recent 
96: experiment\cite{cava2} which shows superconductivity in a 
97: rather narrow range 
98: of doping $\frac{1}{4}<x< \frac{1}{3}$, than predicted by RVB theories, 
99: suggest that there are perhaps left out interactions and consequent 
100: competing phases which make a simple t-J model valid only for 
101: limited range of $x$. The situation is not unusual - even in cuprates 
102: a simple t-J modeling is strictly valid only in the neighborhood of 
103: optimal doping. Charge order phenomenon\cite{stripe,stripe1} is known 
104: in cuprate superconductors; and it has been suggested to  
105: compete\cite{gbstripe} with superconductivity.
106: 
107: Doped \cob, compared to $CuO_2$ layers of high \tc cuprates, has a 
108: narrower conduction band and a less polarizable valence band of oxygen. 
109: Consequently, short range coulomb repulsions among carriers  
110: are screened less. This is likely to stabilize a variety of frustrated 
111: charge ordering in the triangular lattice, as we discuss in this 
112: paper. NMR result of Ray et al.\cite{susc} indeed provides a first evidence 
113: for charge freezing in \naxcob family ($x = \frac{1}{2}$), \ncob below 
114: about $T \approx 300 K$, a large temperature scale. 
115: 
116: We estimate and include unscreened short range coulomb interactions 
117: in the t-J model for the study of unhydrated \naxcob. 
118: We show that in addition to superconductivity, charge ordering in the 
119: narrow conduction band of \cob layer is a major instability, for a 
120: range of higher doping than suspected (Wen et al. in ref 7). 
121: As the unscreened
122: near neighbor coulomb interactions are large and comparable to the
123: band width, the characteristic charge order temperatures are
124: $T_{\rm ch} \sim 400 K$. 
125: 
126: Fortunately, \wat, in hydrated \nxcob, makes t-J modeling valid 
127: for a range of doping. For reasons which we elaborate in the 
128: present paper, hydrogen bonded \wat dipoles of the ice layers screen and 
129: frustrate {\em charge order condensation energy}. That is,
130: they {\em effectively 
131: screen out short range repulsions}, and enable the physics of a simple 
132: t-J model to be realized in a narrow range of doping, 
133: $\frac{1}{4}<x< \frac{1}{3}$, as superconductivity. 
134: 
135: We discuss few important charge ordered states at commensurate 
136: fillings, $x = \frac{1}{4}$ and $ \frac{1}{3}$, which we believe are
137: competitors to the experimentally observed superconductivity, in the 
138: range $\frac{1}{4} < x < \frac{1}{3}$. These reference charge ordered 
139: states are strongly frustrated by the random potential from the 
140: neighboring $Na$ layers, resulting in a glassy phase in the region
141: $\frac{1}{4} < x < \frac{1}{3}$ and beyond. The charge glass phase
142: is likely to be an anomalous metal, very much like the spin gap phase
143: in cuprates, where there are local charge order activities
144: at low frequency scales. 
145: 
146: We estimate the enhancement of the background static dielectric 
147: constant at short distance due to hydrogen bonding in the \wat layers.
148: We find that this screening is sufficient to reduce the large charge
149: order transition temperature down to $\sim 1 K$ and allow 
150: superconductivity to emerge. Strong commensurability effects and
151: the associated short range charge order reduce superconducting \tc 
152: considerably as we approach the commensurate ends $ x = \frac{1}{4} $
153: and $\frac{1}{3}$.
154: 
155: As \wat stabilize superconductivity and discourages charge
156: order, we have a new access to the electronic phases of \cob layer
157: through water. This leads to the possibility of some novel effects: i) 
158: `Electrical Modulation of Superconductivity' 
159: by external electric field or microwave radiation and 
160: ii) `Electroresistance Effect' in the normal state.
161: We estimate that voltages $\sim 500 V$, applied capacitively 
162: to thin films of \nxcob of thickness $\sim~1$ micron will 
163: orient the water dipoles and reduce the short distance 
164: dielectric screening, resulting in stabilization of charge glass order
165: phase and destabilization of superconductivity. This interesting 
166: switching effect may have device potential.
167: 
168: Recently we modeled the low energy physics of doped \cob using 
169: a t-J model and discussed an RVB scenario for superconductivity including
170: a PT violating $d_1+id_2$ wave superconductivity and a $p_1+ip_2$ wave 
171: superconductivity at a higher doping. To study charge order we must 
172: include some leading short distance carrier-carrier and carrier-$Na$-ion 
173: screened coulomb interaction: 
174: \bearr
175: H_{{\rm tJV}} =
176:  -t  \sum_{\langle ij\rangle} C^{\dagger}_{i\sigma} C^{}_{j\sigma}
177: + H.c. +  J \sum_{\langle ij\rangle} 
178: ({\bf S}_{i }\cdot{\bf S}_{j} - \frac{1}{4}n_i n_j) \nonumber \\
179: \sum_{ij} V_{ij} (n_i-1)( n_j-1) + \sum_{i}\epsilon_i (n_i -1)
180: \eearr
181: Here $C$'s and ${\bf S}$'s are the electron and spin operators.
182: As we have an electron doped system we have the `zero occupancy' 
183: constraint $\sum_{\sigma} n^{}_{i\sigma} \neq 0 $ at every site i. 
184: 
185: Recall that doubly occupied $Co^{3+}$ sites carry a charge 
186: $-e$ with reference to the neutral \cob layer and $V_{ij}$ is the 
187: screened coulomb repulsion between them. We have ignored the small two body 
188: off-diagonal coulomb interaction terms. For practical purposes only 
189: the nearest and next nearest neighbor terms $ V_1 \approx  
190: \frac{e^2}{\varepsilon_{ab} R_{nn}} e^{-\frac{R_{nn}}{\lambda_{ab}}}$
191: and $ V_2 \approx  \frac{e^2}{\varepsilon_{ab} R_{nnn}}
192: e^{-\frac{R_{nnn}}{\lambda_{ab}}}$ are important.
193: Here $\varepsilon_{ab} \approx \varepsilon_O + \varepsilon_{\rm {H_2O}} $ 
194: represents the short distance dielectric screening arising from the filled 
195: oxygen bands of \cob layers and $H_2O$ layers in \nxcob. And 
196: $\lambda_{ab} \approx Co-Co$ distance is the Thomas Fermi screening length 
197: for our tight binding metallic layer. The random site energy $\epsilon_i$ 
198: of charge degree of freedom represents the screened 
199: coulomb attraction from neighboring $Na^+$ ions.
200: 
201: Electronic structure calculations\cite{singh} give a value of 
202: $t \approx -0.1~eV$ for the conduction band of the \cob layer. 
203: We estimate $V_1$ and $V_2$ for \naxcob, 
204: the non-hydrated case. The dielectric constants of oxides of $Fe$ and $Ni$ 
205: that flank $Co$ in the periodic table are $\sim 4$ to $12$. We assume a 
206: background (short distance) static dielectric constant of 
207: $\varepsilon_{ab} \approx 8$ for our \cob layer. Recall that in 
208: cuprates the background $\epsilon$ is large $\sim 20$, in view of 
209: a more polarizable octahedral oxygen network; in \cob the oxygen filled
210: band is less polarizable and relatively deep below the fermi level.
211: Using this 
212: dielectric constant and values of Co-Co distances in \ncob 
213: we get $V_1 \approx 0.8~eV$ and $V_2 \approx 0.4~eV$. The mean square
214: fluctuation of the carrier site energy due to disordered $Na$-ions
215: is ${\sqrt{\langle \delta \epsilon_i^2 \rangle}} \approx 0.2~eV $ 
216: 
217: In the absence of hopping, the dopant carriers $Co^{3+}$ will order 
218: classically and undergo order-disorder transition at a fairly high 
219: temperature $k_B T_{\rm ch}(\rm classical)
220: \approx 2{\bar V}\sim 10^3~K$; here 
221: ${\bar V} \equiv \frac{1}{2}(V_1 + V_2)$ is a mean short distance 
222: repulsion. However the electron dynamics reduce 
223: $T_{\rm ch}(\rm classical)$ considerably. To estimate this reduction we
224: perform a mean field analysis of the t-J-V model for a CDW order, 
225: pretending that an unfrustrated charge order arises from nesting 
226: instability for \ncob. This gives us a BCS like expression for \tc:
227: \be
228: k_BT_{\em ch}  
229: \approx \epsilon_F e^{-\frac{1}{{\bar V}\rho_o}} 
230: \ee
231: Here $\rho_o$ is
232: a fermi sea averaged particle-hole density of states corresponding to 
233: the ordering wave vector. Substituting $\epsilon_F \approx 0.5~eV$, 
234: $\rho_0 \approx \frac{1}{2\epsilon_F}$ and ${\bar V} \approx 0.4$, we get  
235:  $T_{\rm ch} \approx 480 K$. Frustration on the triangular lattice at 
236:  half filling
237: and disorder effect from $Na$ ions will further reduce this. Thus we get
238: a charge order temperature in the right range, 
239: $T_{\rm ch}(NMR) \approx 300~K$, seen in NMR, 
240: 
241: \begin{figure}[h]
242: \epsfxsize 5.5cm
243: \centerline {\epsfbox{watfig1.eps}}
244: \caption{`Classical' charge order for $x = \frac{1}{4}$.
245: Spin-0, charge $-e$ carriers ($Co^{3+}$) form a triangular lattice. 
246: Neutral sites with spin-\half moments ($Co^{4+}$) form a Kagome lattice. 
247: Quantum fluctuations will reduce 
248: the amplitude of charge order substantially. Accompanying charge order, 
249: we expect some interesting spin liquid phase or complex short range spin 
250: order at low temperatures}
251: \end{figure}
252: Before we consider influence of \wat we discuss some simple charge 
253: orders at $x = \frac{1}{4}$ and $\frac{1}{3}$ that are favored by 
254: electrostatics in the unhydrated \naxcob. 
255: We ignore the superexchange contribution, as $J <<  V_1, V_2$. 
256: For $x= \frac{1}{4}$, the $Co^{3+}$ sites are arranged on a triangular
257: lattice (figure 1) to minimize coulomb repulsions. Interestingly, 
258: the remaining sites carry spins and form spin-\half Heisenberg
259: antiferromagnet on a Kagome lattice.  In our convention, the classical 
260: energy of this state is $E_{\frac{1}{4}}(\rm Kagome) = 0$. 
261: In the real system, carrier delocalization will considerably 
262: reduce the amplitude of charge order. {\em In this sense the charge ordered
263: states shown in figure 1 and figure 2 are to be thought of as reference
264: classical states}.
265: 
266: Another ground state comparable in energy is an anisotropic metal. It has 
267: ordered stripes - alternating insulating and 0.5 electron doped chains. 
268: The electrostatic energy of this state per site is $ \frac{1}{4}(V_1 + V_2)$.
269: However, the carrier delocalization in the 0.5 electron doped chains leads 
270: to a gain in kinetic energy which is easily estimated when J is 
271: neglected in our t-J model. This case corresponds to a quarter filled 
272: infinite U Hubbard model, which can be converted into a half filled band of 
273: non-interacting spinless fermions giving us the delocalization energy
274: $ = -|t|\sum \cos k = -2{\frac{|t|}{\pi}}$. Thus we get a total energy
275: per site, 
276: $E_{\frac{1}{4}}({\rm stripe}) = \frac{1}{4}(V_1 + V_2) - 
277: 2{\frac{|t|}{\pi}}$.  
278: 
279: Figure 2 shows the case of $x = \frac{1}{3}$. This classical 
280: ground state minimizes electrostatic repulsion and $Co^{3+}$ sites fill
281: one of the three sublattices and the remaining hexagonal lattice is
282: the neutral spin-\half site. It is a hexagonal spin-\half quantum 
283: antiferromagnet. The energy of this state per site is $E_{\frac{1}{3}} = 
284: \frac{V_2}{2}$. We also find striped states which are local minima.
285: \begin{figure}[h]
286: \epsfxsize 5.5cm
287: \centerline {\epsfbox{watfig2.eps}}
288: \caption{`Classical' charge order for $x = \frac{1}{3}$.
289: Triangular lattice of localized charge $-e$ carriers and a 
290: Hexagonal lattice of neutral spin-\half moments. Quantum fluctuations
291: will reduce the amplitude of charge order substantially. Accompanying
292: charge order we expect short range AFM order at low temperatures}
293: \end{figure}
294: 
295: So far we studied charge order in the \cob plane at commensurate
296: fillings. As we move away from $ x = \frac{1}{3}$ and $ x = \frac{1}{4}$,
297: defects and discommensurations will be produced or we may go to an 
298: incommensurate charge ordered structure. There may be one or more 
299: first order phase boundary between $x = \frac{1}{3}$ and 
300: $ x = \frac{1}{4}$. However, all these nice charge ordered phases
301: will be challenged by a generically disordered arrangement of \na 
302: ions in a triangular lattice and the consequent random potential seen
303: by the mobile $Co^{3+}$ carriers as explained below . 
304: 
305: The energetically preferred sites of \na atoms\cite{cobstr}
306: in \naxcob form a triangular lattice that have 
307: the same lattice parameter as the triangular $Co$ layer. However,
308: the \na and $Co$ lattices are relatively shifted - if we project
309: the allowed positions of the \na atoms of the nearest top and 
310: bottom layer onto the $Co$ layer, these sites become the dual lattice
311: (hexagonal lattice) of the $Co$ triangular lattice. Because of this
312: {\em a sublattice order of the \na atoms does not couple to the
313: charge density wave order-parameter of the \cob lattice and in 
314: principle allows a finite temperature charge order-disorder 
315: phase transition}:
316: a 3-state Potts ($Z_3$ symmetry) model transition at $x = \frac{1}{3}$ 
317: and a 4-state Potts model ($Z_4$ symmetry) transition at 
318: $ x = \frac{1}{4}$. However, an inevitable  disorder in \na 
319: sublattice leads to, based on an Imry-Ma  type of argument, a glassy 
320: order at low temperatures rather than a genuine charge order phase 
321: transition. Thus we expect a phase diagram depicted in figure 3 for 
322: the range $\frac{1}{3} < x < \frac{1}{4}$. 
323: 
324: Let us move on to the hydrated case, \nxcob. As we mentioned 
325: earlier, the enhanced dielectric constant of the \wat layer
326: will screen short range coulomb repulsion and weaken and melt 
327: the high temperature charge ordering. It also screens and weakens
328: the random \na potential seen by the carriers.
329: For the appearance of low temperature superconductivity 
330: $T_{\rm ch}$ need not be reduced to nearly zero value. A 
331: sufficiently weakened charge ordered state may give up at low 
332: temperatures and superconductivity may emerge. Figure 3
333: shows sketches the change of phase diagram as we go to the
334: hydrated case. A strong resistance anomaly seen in a recent 
335: experiment\cite{jin}  at $T^* \approx 50~K$, may be 
336: the weakened charge order transition that we are discussing.
337: \begin{figure}[h]
338: \epsfxsize 8.0cm
339: \centerline {\epsfbox{watfig3.eps}}
340: \caption{Schematic phase diagram showing how hydration affects 
341: electronic phases: superconductivity is stabilized at low 
342: temperatures by a strong suppression of charge glass order.}
343: \end{figure}
344: Let us discuss nature of \wat ordering hydrogen bonding 
345: in \nxcob in some detail. In what follows we use a recent 
346: suggestion of Cava et al.\cite{cava3} that \wat may have a structure 
347: similar to the layers in hexagonal ice (1h ice). As mentioned earlier,
348: in \naxcob, energetically favorable interlayer sites of $Na$ form 
349: a triangular lattice. These sites are at the center of trigonal 
350: biprisms, capped by an oxygen atom at the top and one at the bottom. 
351: To understand \wat ordering, we consider
352: $Na_{\frac{1}{3}}.CoO_2.{\frac{4}{3}}H_2O$ as a reference compound.
353: Fill one of of the three sublattices by \na atoms to minimize 
354: electrostatic energy.  We are left with two empty sublattices that
355: form a hexagonal lattice. Two \wat molecules may be accommodated at the 
356: top and bottom of the capped trigonal biprism. By doing so we get two
357: hexagonal lattices of \wat sandwiching a triangular lattice of \na
358: ions. Thus we have a triangular lattice filled by \na and \wat in the
359: ratio $1:4$. We can view the above hexagonal sheets as the sheets in
360: hexagonal ice structure, as suggested by Cava and collaborators\cite{cava3}
361: in their
362: preliminary studies. {\em We find it very interesting that the nearest
363: neighbor $H_2O$-$H_2O$ distance in the above geometry, $ \approx 2.81~Au$,
364: is nearly the same\cite{icebook}  as that in real hexagonal ice 
365: $\approx 2.71$}.
366: No wonder water may freeze into ice in \nxcob ! Having nearly the same
367: \wat-\wat distance may also help ice sheets to have a good hydrogen 
368: bonding network like in hexagonal ice.
369: 
370: Thus it is likely that \wat molecules continue to have hydrogen bonding
371: in spite of the \cob and \na environment. As hydrogen bond energy is
372: substantial $\sim 0.5~ eV$, water tends to have hydrogen bonding 
373: activity in extreme environments. Examples are biological systems, 
374: clathrate hydrates and water containing charged ions, where \wat 
375: continue to maintain hydrogen bonding even though the 
376: local structure deviates considerably from the standard ice or water 
377: structure. Further the random \na environment in \nxcob may convert
378: the 2D ice layer into a 2D amorphous ice layer with a good short range 
379: hexagonal order.
380: 
381: Now we discuss how the dielectric property of the \wat layer may control
382: the low temperature electronic phases of the conducting \cob layer. We
383: are interested in finding how the electron-electron interactions 
384: at the charge order wave vector ${\bf q } = {\bf Q}$ 
385: get screened by the interacting water dipoles. The relevant static 
386: dielectric constant is $\epsilon_{\rm H_2O} (Q)$. In an ice system like 
387: ours with a large disorder in dipole orientations, we expect very small 
388: variation of $\epsilon_{\rm H_2O}(q)$ with $q$. Further random site 
389: disorder in the 
390: \na sublattice will produce Bjerrum defects; so we do not 
391: expect a 2D dipolar order-disorder transition, as in ideal 2D models 
392: of ice. 
393: 
394: Dielectric constant of ice has been studied extensively in the past 
395: and also recently\cite{frohlich,haymet,icebook}. A general expression 
396: for the dielectric 
397: constant of an interacting dipolar system is\cite{frohlich,haymet}:
398: \be
399: \epsilon = \epsilon_{\infty} + 
400: \frac{4\pi}{3Vk_BT} \langle ({\bf P} - \langle {\bf P}\rangle )^2 \rangle
401: \ee
402: Here $\epsilon_{\infty} \sim 1 - 2$ is the high frequency dielectric 
403: constant of the dipole, in our case \wat molecule. $\bf P$
404: is the total dipole moment of the system of volume V. And 
405: $\langle~...~\rangle$ denotes thermal average.  Static dielectric
406: constant of ice has not been measured at liquid He temperatures, as the
407: dielectric relaxation becomes too slow even around liquid air 
408: temperatures. Fortunately, extensive numerical study of \ewat
409: are available. For example, a recent calculation\cite{haymet} shows 
410: that for hexagonal ice, \ewat $\approx 220 $ at $T = 50 K$. 
411: 
412: We use this 3D result to get an approximate estimate for our weakly 
413: coupled hexagonal ice layers as follows.  We replace the volume 
414: $V$ by $\approx 6V$ to account for the c-axis expansion
415: in \nxcob. Missing hydrogen bonds along the c-axis reduces the number
416: of allowed proton configurations leading to a reduction of 
417: $\langle ({\bf P} - \langle {\bf P}\rangle )^2 \rangle$ to  
418: $\approx \frac{2}{3} \langle ({\bf P} - \langle {\bf P}\rangle )^2 \rangle$.
419: This gives us \ewat (hexagonal sheet) $\approx 20$ at $T = 50 K$. Our
420: system being strongly disordered, we do not expect \ewat (hexagonal sheet)
421: to change at lower temperatures. Thus the background dielectric constant
422: of \nxcob is $\epsilon = \epsilon_o + \epsilon_{\rm H_2O} \approx
423: 8 + 20$. This reduces the mean short range repulsion by nearly a factor
424: of 3, making the $T_{\rm ch} \approx 1 K$, in equation (2). Once the 
425: long range charge order
426: is disabled by a reduction of V, the simple t-J model and consequent
427: low temperature superconducting phase is realized, albeit with a reduced
428: \tc in the range $\frac{1}{3} < x < \frac{1}{4}$. The sharp reduction
429: in superconducting \tc at the commensurate boundaries of the dome should 
430: arise from the strong short range order and lesser discommensurations and
431: defects. The experimentally seen flat value of \tc $\approx 2 K$ 
432: for $x < \frac{1}{4}$ and for $x > \frac{1}{3}$is likely to be an effect 
433: of phase separation.
434: \begin{figure}[h]
435: \epsfxsize 6.0cm
436: \centerline {\epsfbox{watfig4.eps}}
437: \caption{Schematic experiment to observe the `Electrical  
438: Modulation of Superconducitivy'. The order voltage required
439: is $500~V$, when the film thickness is about a micron.}
440: \end{figure}
441: Our proposal of a critical and catalytic role of \wat layer suggests
442: ways to access and control low temperature electronic phases of the 
443: \cob layer. Based on this we suggest two effects: i) `Electrical 
444: Modulation of Superconductivity'. Here we control (figure 4) 
445: superconductivity 
446: and superconducting \tc by modifying the screening property of \wat
447: layer by external electric fields - DC, AC or pulsed fields. 
448: This has interesting consequences of 
449: being able to locally erase superconductivity by STM tips, dynamically
450: create Josephson networks or create 2D superconductivity of desired
451: shapes through appropriate capacitor shapes etc. 
452: Since microwaves are absorved by hydrogen bonded networks we can pump 
453: microwaves at appropriate frequencies ($\hbar \omega <
454: \Delta_{\rm sc}$, the superconducting gap) and dynamically polarize 
455: water dipoles and may influence its dielectric properties, and in turn 
456: control superconductivity.
457: 
458: ii) `Electroresistance Effect'. By modifying the amplitude of 
459: charge glass order as well as $T_{\rm ch}$ in the 
460: non-superconducting state, by influencing the \wat layer by
461: external DC or AC electric field,
462: we can change $\rho_{ab}$, the ab-plane resistivity. 
463: 
464: Below we estimate the electric field required to completely suppress 
465: superconductivity. Having established a hydrogen bonded
466: network it requires a finite energy to rotate a water molecule and
467: orient its dipole moment along an external electric field. In infrared 
468: absorption
469: and neutron scattering experiments\cite{icebook}
470: the absorption band corresponding to 
471: rotation of \wat molecules is in the range $60~{\rm to}~120~meV$. 
472: Assuming a random orientation, the average energy required to reorient 
473: a water molecule is $\approx 50~meV$, i.e., a potential of
474: $50~{\rm mV}$ applied over the length $\approx 1~Au$ of the
475: water dipole will orient the dipole moment of water along its field.
476: If we have c-axis oriented \nxcob film of thickness $1$ micron we need
477: to apply a voltage $\approx 500~Volts$ across the film, in order to 
478: orient the majority of dipoles. Strong polarization of water dipoles 
479: reduces the dielectric constant, as is evident from equation (3).
480: The resulting reduced screening of carriers in the \cob layer
481: allows charge order to grow and superconductivity gets suppressed.
482: 
483: A theoretical analysis, including some of the possible difficulties 
484: in observing the effects will be presented in a future publication. 
485: 
486: To get a clear understanding of this complex system, and to see
487: if our proposal is correct more experiments are necessary:
488: 
489: a) {\bf Charge Order:} It will be interesting to perform NMR,
490: NQR, STM, $\mu SR$ and other local probe measurements to look for 
491: charge order in the vicinity of the commensurate fillings 
492: $x = \frac{1}{4}, \frac{1}{3}, \frac{1}{2}, \frac{2}{3}$ and 
493: $\frac{3}{4}$ and see how they differ between the two systems
494: \nxcob and \naxcob. 
495: 
496: b) {\bf Spin order, singlets and gaps:} Accompanying local charge order 
497: we expect a spin order at 
498: low temperatures (the scale of J is small, $\sim 6$ to $7~meV$). 
499: In general the enhanced singlet stabilization by the superexchange
500: process will introduce some kind of spin gap phenomenon. If the charge 
501: order at $x = \frac{1}{4}$ leads to a Kagome lattice of spins it will be 
502: an interesting testing ground for some of the ideas of the spin
503: liquid phase of spin-\half Kagome antiferromagnet, including possible
504: novel excitations.
505: 
506: c) {\bf Lower Doping:} Experimentally, it has not been 
507: possible\cite{cava2}
508: to make \nxcob for $x < \frac{1}{4}$. It is likely\cite{cava2} that 
509: c-axis ionic bonding is weakened, by decreasing $x$ and 
510: presence of water layer, making a 3D structure unstable.
511: It will be important to synthesize, by 
512: non-equilibrium means, meta stable compounds for $x < \frac{1}{4}$
513: to test the validity of RVB theory and also test our hypothesis
514: of the role played by water.
515: 
516: d) {\bf Replacing \wat:} It will be desirable to have a stable solid
517: $Na_x CoO_2.y X$, where an intercalant `X' not only increases the 
518: dielectric constant but also provides additional bonding between 
519: \cob layers and make stable compounds for $x < \frac{1}{4}$.
520: 
521: e) {\bf Higher doping:} According to reference 6, the dopant 
522: induced dynamics, within the t-J model will favor ferromagnetic 
523: correlations and a consequent p-wave superconductivity at higher 
524: dopings slightly above $x = \frac{1}{3}$. It will be interesting 
525: to look for this.
526: 
527: f) {\bf Inhomogeneous Superconductivity:} If ice plays a central
528: role, as suggested in this paper, \wat density fluctuation in  
529: ice layer will directly influence superconductivity in nearby 
530: \cob layers resulting in a corresponding fluctuation in the 
531: superconducting order parameter and possible well grown local charge 
532: ordered phase. 
533: 
534: g) {\bf Slow Relaxation:} Since interacting water dipoles have very 
535: slow dielectric relaxation time scales\cite{icebook}, they may 
536: consequently affect superconductivity and impose some anomalous 
537: relaxation/aging effects. 
538: 
539: The present paper is phenomenological and qualitative in character.
540: Any detailed quantitative calculations of \tc and phase diagram
541: for this complex system needs further experimental guidance.
542: Issue of calculating local screening and dielectric constant in 
543: hydrogen bonded systems is known to have subtleties; added to 
544: this, we have conducting layers sandwiching water layers. We have 
545: made very crude estimates based on simple physical arguments and
546: very approximate considerations, as our primary aim is to focus 
547: and identify how water could play a crucial role in this complex 
548: system.
549: 
550: We thank A.K. Mishra, V.N. Muthukumar, Debanand Sa, Manas Sardar 
551: and R. Shankar for discussion and Latha Malar Baskaran for 
552: reference [18].
553: 
554: \begin{references}
555: \bibitem{takada} K. Takada et al., Nature, {\bf 422} 53 (03)
556: \bibitem{cava1} M. Foo et al., cond-mat/0304464
557: \bibitem{cava2} Y. Wang et al., cond-mat/0305455
558: \bibitem{chu} B. Lorenz et al., cond-mat/0304537;
559: F. Rivadulla et al., cond-mat/0304455
560: \bibitem{jin} R. Jin et al., cond-mat/0306066
561: \bibitem{gbcob} G. Baskaran, cond-mat/0303649
562: \bibitem{rvbcob}Brijesh Kumar, B.S. Shastry, cond-mat/0304210;
563: Qiang-Hua Wang, Dung-Hai Lee and Patrick A. Lee, cond-mat/0304377;
564: Masao Ogata, cond-mat/0304405
565: \bibitem{pwagbrvb} P.W. Anderson, Science, {\bf 235} 1196 (87)
566: G. Baskaran, Z. Zou and P.W. Anderson, 
567: Sol. St. Commn, {\bf 63} 973 (87); G. Baskaran and P.W. Anderson, 
568: Phys. Rev. {\bf B~37} 580 (88)
569: \bibitem{stripe} J. Zaanen and O. Gunnarsson, Phys. Rev.
570: {\bf B~46} 7391 (89); V.J. Emery, S. A. Kivelson and
571: O. Zachar, Phys. Rev. {\bf B~56} 6120 (97); S.R. White and
572: D.J. Scalapino, Phys. Rev. {\bf B~60} R753 (99);
573: \bibitem{stripe1} J.M. Tranquada et al., Nature {\bf 375} 561 (95)
574: \bibitem{gbstripe} G. Baskaran, Mod. Phys. Lett., {\bf B~14} 377 (00)
575: \bibitem{susc} R. Ray et al., Phys. Rev. {\bf B~59} 9454 (99);
576: Phys. stat. sol. {\bf B~215} 703 (99)
577: \bibitem{singh} D.J. Singh, Phys. Rev. {\bf B~61} 13397 (00)
578: \bibitem{cobstr} R.J. Balsys and R.L. Davis, Sol. St. Ionics,
579: {\bf 93} 279 (96)
580: \bibitem{cava3} R. Cava, M2S-Rio Meeting, 25-30 May 2003.
581: \bibitem{icebook} V. F. Petrenko and R. W. Whitworth, 
582: Physics of Ice, (Oxford University Press, 1999)
583: \bibitem{frohlich} H. Frohlich, Theory of Dielectrics and
584: Dielectric Loss (Oxford University Press, Oxford, 1958)
585: \bibitem{haymet} S.W. Rick and A.D.J. Haymet, 
586: J. Chem. Phys. {\bf 118} 9291 (2003)
587: \end{references}
588: \end{multicols}
589: \end{document}
590: 
591: