Vous êtes sur la page 1sur 39

Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 1

PREPARATION AND THERMOPOWER BEHAVIOR OF


THE NEW COBALT OXYHYDRATES
(NA,K)X(H2O)YCOO2 PREPARED BY THE AQUEOUS
PERMANGANATE SOLUTION

Chia-Jyi Liu
Department of Physics, National Changhua University of Education, Changhua 500,
Taiwan, R. O. C.

ABSTRACT

This chapter would present a novel route to prepare the superconductive sodium
cobalt oxyhydrates using aqueous permanganate solution. Immersing γ-Na0.7CoO2 in
KMnO4 (aq.) leads to two distinct novel phases of (Na,K)x(H2O)yCoO2, depending on the
concentration of the KMnO4 (aq.) solution. These two phases differ in the c-axis with little
change from the parent compound in the a-axis. The phase transformation can be viewed
as a topotactic process. The c ≈ 19.6Å phase has bi-layers of water molecule inserted into
the lattice, whereas the c ≈ 13.9Å phase has mono-layers of water molecule inserted into
the lattice essentially without changing the skeleton of the parent compound. Immersing
γ-Na0.7CoO2 in NaMnO4 (aq.) only leads to the c ≈ 19.6Å phase of Nax(H2O)yCoO2. The
formation mechanism and phase stability of KMnO4-treated (Na,K)x(H2O)yCoO2 would
be discussed. The hydration is a very slow process when using KMnO4(aq.) with a low
molar ratio of KMnO4/Na. Three related phases of c ≈ 19.6Å, c ≈ 13.9Å, and c ≈ 11.2Å
exist in the cobalt oxyhydrates, being associated with the water content in the lattice. The
thermal stability of the c ≈ 19.6Å phase of (Na,K)x(H2O)yCoO2 behaves differently from
that of the c ≈ 13.9Å phase, which is ascribed to the difference between the ion-dipole
interaction of K+-H2O within the alkaline layers and that of Na+-H2O located between the
CoO2 layers and the alkaline layers. The thermopower behavior of three related phases of
cobalt oxyhydrates will be presented. The generalized Heikes formula would be used to
explain the variation of the thermopower between γ-Na0.7CoO2 and
Na0.33K0.02(H2O)1.33CoO2 arising from the variation of the concentration of Co4+.
2 Chia-Jyi Liu

1. INTRODUCTION
The discovery of superconductivity of the cobalt oxyhydrate Na0.35(H2O)1.3CoO2 with Tc
≈ 4-5 K by Takada et al. [1] is a surprise after several years of studies on its parent
compounds of sodium cobalt oxides γ-NaxCoO2. The γ-Na0.5CoO2 is a potential candidate for
thermoelectric applications due to its high electrical conductivity, large thermopower and low
thermal conductivity. [2] It is an amazing and easy procedure to convert thermoelectric
material to superconductor through virtually a one-pot reaction by immersing the parent
material γ-Na0.7CoO2 in Br2/CH3CN solution for 5 days followed by washing with water. The
structure of both the γ-Na0.7CoO2 and Na0.35(H2O)1.3CoO2 belongs to a hexagonal crystal
system with the space group of P63/mmc (No. 194). For γ-Na0.7CoO2, there are two CoO2
layers of edge-sharing CoO6 octahedra in the unit cell with the Na ions sandwiched between
the two CoO2 layers. The Na ions of γ-Na0.7CoO2 are located in two different crystallographic
sites, i.e., 2b and 2d, and tend to disassociate from the crystal lattice when immersed in
aqueous or acetonitrile solution. After intercalating water molecules into the lattice of parent
material γ-Na0.7CoO2, the c-axis elongates from ca. 10.9 Å to ca. 19.6 Å, while the a-axis
shows little change. For Na0.35(H2O)1.3CoO2, some of the water molecules are located between
the CoO2 layers and the Na ion layers, and others are located within the Na ion layers. The
former tends to escape from the crystal lattice in the ambient environment accompanied by a
shrinkage of the c-axis to ca. 13.8 Å with little change of the a-axis; the latter would
dissociate from the lattice only upon heating due to a stronger bonding of water molecules
with the Na ions within the Na ion layers as compared to the former. Complete removal of the
water molecules would result in the de-hydrated form of Na0.35CoO2 with further shrinkage of
the c-axis to ca. 11.2 Å, which is a little bit longer than c ≈ 10.9 Å for the parent material γ-
Na0.7CoO2.
Due to the nature of the hexagonal lattice, the arrangement of Co ions forms a triangular
lattice. As a result, it would frustrate the spin arrangement of electrons in such a triangular
lattice if requiring the spin to align antiparallel to each other. It is a common way to ease up
the electron-electron repulsion for an electron to align its spin antiparallel to its nearest
neighbor for a system with strong electron correlations. Based on the heat capacity, [3]
magnetic, [4] thermopower, [5] and angle-resolved photoemission spectroscopy (ARPES) [6]
measurements, it has been suggested that NaxCoO2 is a system with strong correlations of 3d
electrons. It would be interesting to correlate the frustrated spin arrangements or the spin state
with the superconductivity. Na0.35(H2O)1.3CoO2 is not only one of the few examples of
layered transition metal oxide superconductors without containing copper, it is but also a
particularly interesting system for comparison with the high-Tc cuprates in terms of the
structure-electronic state correlations in light of the fact that both have 2D layers (triangular
CoO2 layers and square CuO2 layers) in structure and have spin 1/2 ions (t2g5 for Co4+ in low
spin state and t2g6eg3 for Cu2+) in electron configuration. In spite of their structural difference,
mixed valency is a common feature to both systems. Besides, the family of sodium cobalt
oxides NaxCoO2 is of particular interest because of their structural, magnetic and
thermoelectric properties.
This chapter is organized as follows. Section 2 describes the crystal structures of the
sodium cobalt oxides NaxCoO2. Section 3 discusses the preparation methods, phase formation
mechanism of the superconducting cobalt oxyhydrates and related compounds using the
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 3

aqueous permanganate solutions, and the kinetics of the topotactic transformation.. Section 4
discusses the phase stability of the permanganate-treated cobalt oxyhydrates
(Na,K)x(H2O)yCoO2. Section 5 describes the magnetization measurements on
(Na,K)x(H2O)yCoO2 and discusses the thermopower behavior of three related phases of cobalt
oxyhydrates. The generalized Heikes formula is used to explain the variation of the
thermopower due to the variation of the spin and orbital degeneracy in the 3d electrons of the
cobalt ions.

2. CRYSTAL STRUCTURES OF SODIUM COBALT OXIDES


NAXCOO2 WITH X ≤ 1
According to Fouassier et al., there are four bronze type phases for NaxCoO2 with x ≤ 1,
where the phase formation is associated with the value x of the sodium content, i.e., (1) α-
NaxCoO2 (0.9 ≤ x ≤ 1); (2) α'-Na0.75CoO2; (3) β-NaxCoO2 (0.55 ≤ x ≤ 0.6); and (4) γ-
NaxCoyO2 (0.55 ≤ x/y ≤ 0.74). [7] Fig. 1 shows the crystal structures of α-, β-, and γ-phase of
NaxCoO2. For all the four phases, the volatile sodium ions are located between CoO2 layers
and are considered as a charge reservoir. Hence the valence state of Co ions can be tuned by
varying the Na content in the materials, which would have significant effects on their
structural, electrical, magnetic, and thermoelectric properties. The α-NaxCoO2 is isomorphous
to α-NaFeO2 and has a trigonal lattice with the space group R3 m (No. 166). There are three
crystallographically distinct positions: Co at 3b (0,0,0.5), Na at 3a (0,0,0), and O at 6c (0,0,z).
[7,8] The α'-NaxCoO2 is isomorphous to α-NaMnO2 and has a trigonal lattice with the space
group C2/m (No. 12). There are three crystallographically distinct positions: Co at 2a (0,0,0),
Na at 2d (0,0.5,0.5), and O at 4i (x,y,0). [7,9] Fouassier et al. considered the crystal structure
of β-NaxCoO2 to have a trigonal lattice with the space group of R3m (No. 160), but the recent
studies show that the structure of β-NaxCoO2 should be assigned to have a monoclinic lattice
with the space group C2/m (No. 12). There are three crystallographically distinct positions:
Co at 2a (0, 0, 0), Na at 8j (0.810, 0.097, 0.495), and O at 4i (0.3898, 0, 0.1790). [7,10,11]
The γ-NaxCoO2 is isomorphous to α-NaMnO2 and has a hexagonal lattice with the space
group of P63/mmc (No. 194). There are four crystallographically distinct positions: Co at 2a
(0,0,0), Na(1) at 2b (0,0,0.25), Na(2) at 2d (0.333,0.667,0.75), and O at 4f (0.333,0.667,z).
[7,12]
4 Chia-Jyi Liu

(a)

(b)
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 5

(c)
Figure 1. Crystal structures of (a) α phase; (b) β phase; (c) γ phase.(Na: ; Co: ; O: )

The material γ-Na0.7CoO2 belongs to a category of γ-NaxCoyO2 phase with 0.55 ≤ x/y ≤
0.74. The Na ions in the gamma phase partially occupy the two crystallographically distinct
sites 2b and 2d, and prefer occupying the 2d site, since the 2b site is directly between the Co
ions when viewing along the c direction and results in a slightly higher energy state. [13,14]
The calculations based on the density-functional theory within the local density
approximation with the Ceperly-Alder exchange-correlation functional suggest that the
arrangements of Na ions within a given plane of 2b and 2d tend to have ordering structures
driven by the screened electrostatic interactions among the Na ions. [14] The Na ordering
structures have been observed by electron diffraction experiments. [15] As a matter of fact,
the phase formation of NaxCoO2 is not only associated with the sodium content but also with
the firing temperature and partial pressure of oxygen. The γ-NaxCoO2 phase with 0.74 ≤ x ≤ 1
can be readily synthesized by using appropriate firing temperature.

3. MATERIALS SYNTHESIS
Rational approaches to synthesize novel materials have been a great interest to the solid
state chemists for producing solids with interesting properties and potential applications.
Intercalation and ion exchange are the techniques among those low-temperature processes in
the rational approaches. Intercalation refers to the insertion of a guest species into a
crystalline host lattice. Ion exchange has been defined as a process in which ions are released
from an insoluble permanent material in exchange for other ions in a surrounding solution.
Both techniques involve a reversible reaction with the structural integrity of the host lattice
essentially preserved throughout the process. In general, both types of reactions can occur at
relatively low temperatures.
6 Chia-Jyi Liu

3.1 SYNTHESIS RATIONALE


The superconducting phase of fully hydrated Na0.35(H2O)1.3CoO2 was obtained by
immersing γ-Na0.7CoO2 powders in Br2/CH3CN solution followed by filtering and rinsing.
Upon intercalation, the c-axis of the parent compound γ-Na0.7CoO2 expands from 10.9Å to
19.6Å to form the fully hydrated sodium cobalt oxyhydrates Na0.35(H2O)1.3CoO2 with little
change in the a-axis. Since the fully hydrated phase preserves the structural integrity of the
host lattice, the reaction can be viewed as a topotactic transformation process. Formation of
superconducting Na0.35(H2O)1.3CoO2 is also generally considered as an oxidative
deintercalation by removing Na+ partially from the host lattice, followed by a hydration
reaction. It is well-known that KMnO4 is a strong oxidizing agent with a unique affinity for
oxidizing organic compounds, which is particularly useful for in-situ remediation of organic
compounds in ground water and subsurface soils when coupled with delivery techniques.
Superconducting cobalt oxyhydrates can be obtained by immersing γ-Na0.7CoO2 in the
aqueous KMnO4 solution at a low molar ratio of KMnO4 relative to Na in the parent
compound without resort to the volatile and flammable Br2/CH3CN solution. [16] Since
removing more Na+ from the host lattice of γ-Na0.7CoO2 requires a higher oxidation potential,
[17] one would expect that the sodium content in the cobalt oxyhydtates could be tuned by
varying the concentration of aqueous KMnO4 solution due to the fact that a higher
concentration of aqueous KMnO4 solution would have a higher redox potential according to
the Nernst equation. Besides the potassium permanganate, sodium permanganate NaMnO4 is
another common form of permanganates. Both forms have similar chemical reactivity.
Sodium permanganate has about 10 times the solubility in water that potassium permanganate
has. This fact would be an advantage to produce superconducting potassium-free cobalt
oxyhydrates in large quantities using NaMnO4 as the de-intercalation and oxidation agent
without resort to highly toxic Br2/CH3CN solution.

3.2 PREPATION OF (Na,K)x(H2O)yCoO2 AND Nax(H2O)yCoO2


Using the conventional solid state reaction procedure or the sol gel route to prepare
NaxCoO2 normally requires using excess Na to compensate for the loss of the Na during the
gradual heating process. In addition, the loss of Na is accompanied by formation of the
impurity phase of Co3O4. [18,19] In order to avoid the loss of Na, a rapid heat-up procedure
[20] is adopted to prepare polycrystalline parent compounds of γ-Na0.7CoO2. High purity
powders of Na2CO3 and CoO were thoroughly mixed and ground using a Retch MM2000
laboratory mixer mill and calcined in a preheated box furnace at 800°C for 12 h. In case of
using aqueous KMnO4 solution, the resulting powders (0.5 -1 g) were immersed and stirred in
50 - 680 ml of aqueous KMnO4 solution with the molar ratio of KMnO4/Na between 0.05 and
40 at room temperature for 5 days. The concentration of KMnO4 used to prepare
(Na,K)x(H2O)yCoO2 depends on the molar ratio of KMnO4/Na and the amount of water used
to dissolve KMnO4. High molar ratio of KMnO4/Na requires a large amount of water to
dissolve KMnO4 due to the limited solubility of KMnO4 in water. For the 40X sample, it
needs 680 ml water to completely dissolve KMnO4. The resulting products were filtered and
washed by de-ionized water several times with the total volume of 150-200 cc. Note that the
contents of potassium and sodium of the products depend slightly on the volume of water in
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 7

the washing process because they would lose a small amount in the water as confirmed by
ICP analyses. The powders were then stored in a wet chamber with relative humidity of 98%
to avoid loss of the water content. In case of using aqueous NaMnO4 solution, the molar ratio
of NaMnO4/Na is between 2 and 40. The volume of water to rinse off NaMnO4 is 350 cc.

3.3 PHASE FORMATION OF (Na,K)x(H2O)yCoO2 AND Nax(H2O)yCoO2


Figs. 2 and 3 show the powder x-ray diffraction patterns of cobalt oxyhydrates obtained
by immersing γ-Na0.7CoO2 in aqueous KMnO4 solution different molar ratios of KMnO4/Na.
It is clearly seen that

reflection for non-hydrate


Fe Kα
(002)

Na CoO
x 2
0.05X
Intensity (arb. uints)

0.1X
(004)
(002)

(006)

0.3X

0.5X
*
2.29X *

5 10 15 20 25 30 35 40
2θ (degree)

Figure 2. Powder x-ray diffraction patterns of Na0.7CoO2 and (Na,K)x(H2O)yCoO2 obtained by


immersing Na0.7CoO2 in low molar ratio of aqueous KMnO4 solution with respect to the Na content in
the parent compound (0.05 ≤ KMnO4/Na ≤ 2.29). The 0.5X represent the molar ratio of KMnO4/Na is
0.5. The asterisk (*) indicates the (002) reflection of the c ≈ 13.9 Å phase.
8 Chia-Jyi Liu

FeKα
4.286X
Intensity (arb. units)

(108) (110)
(006) (102)
10X
(004)
(002)

(104)
(100)

(112)
(008)
(103)

(106)

20X

40X

10 20 30 40 50 60 70 80 90
2θ (degree)

Figure 3. Powder x-ray diffraction patterns of (Na,K)x(H2O)yCoO2 obtained by immersing γ-


Na0.7CoO2 in high molar ratio of aqueous KMnO4 solution with respect to the Na content in the parent
compound (4.286 ≤ KMnO4/Na ≤ 40).

different molar ratios of KMnO4/Na (different X) result in the variation of XRD patterns. The
XRD patterns can be classified as four categories according to the molar ratio of KMnO4/Na
used to treat the parent material. These four categories are KMnO4/Na≤0.1,
0.3≤KMnO4/Na≤0.4, 0.5≤KMnO4/Na≤2.29, and 4.286≤KMnO4/Na≤40. As shown in Fig. 2,
the characteristic x-ray diffraction peak of the maximum intensity with the Miller index (002)
shifts from 2θ ≈ 20.3° (d spacing ≈ 5.5 Å) for the parent compound irradiated by the Fe Kα
radiation to 2θ ≈ 11.3° (d spacing ≈ 9.8 Å) for the samples between 0.3X and 2.29X,
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 9

indicating that the c-axis expands from c ≈ 10.9 Å to c ≈ 19.6 Å in the unit cell, which is
consistent with those for samples obtained by the Br2/CH3CN solution. Fig. 4 shows the XRD
patterns of the KMnO4-treated 0.05X sample, indicating the transformation of the non-
hydrate phase with c ≈ 11.2 Å to the fully-hydrated phase with c ≈ 19.6 Å. For the 0.05X and
0.1X samples, the XRD patterns are a mixture of a fully hydrated phase (c ≈ 19.6 Å) and a
non-hydrate phase (c ≈ 11.2 Å). However, it is found that when storing the as-prepared 0.05X
and 0.1X samples in a wet chamber with a relative humidity of 98% for a long period of time,
the fully hydrated phase grows at the expense of the non-hydrate phase, as evidenced by the
fact that the (002) reflection peak (2θ ≈ 11.4°) of the fully hydrated phase grows, whereas the
(002) peak (2θ ≈ 20.1°) of the non-hydrate phase nearly disappears. This result indicates that
intercalation of the water molecules could undergo a gas-solid interaction in addition to that
through a liquid–solid reaction.
As shown in Fig. 3, samples synthesized by the high concentration of KMnO4
(4.286≤KMnO4/Na≤40) surprisingly and clearly show distinct XRD patterns from low X
samples. They have the characteristic peak of (002) reflection occurring at 2θ ≈ 16° (d
spacing ≈ 6.95 Å) and resulting in c ≈ 13.9 Å. It has been shown that the fully hydrated phase
obtained by the Br2/CH3CN solution is not stable at ambient conditions and tends to lose
water becoming an intermediate hydrate phase with c ≈ 13.8 Å. [21] Both the XRD pattern
and the size of c-axis of our high X samples are very similar to the intermediate

Fe Kα 0.05X
*

*
as-prepared wet powders
Intensity (arb. units)

after filtration

24 hrs in air after filtration


(002)

(004)

4 days in wet chamber


(006)

5 10 15 20 25 30 35 40
2θ (degree)

Figure 4. Powder x-ray diffraction patterns of the KMnO4-treated 0.05X sample. The as-prepared
wet powders are a mixture of c ≈ 19.6 Å and c ≈ 11.2 Å phases. The asterisk is the (002) reflection peak
of c ≈ 11.2 Å phase. After further exposing the as-prepared powders in the humid surroundings, the c ≈
11.2 Å phase transforms to c ≈ 19.6 Å.
10 Chia-Jyi Liu

hydrate phase with c ≈ 13.8 Å, which could be obtained by heating at ca. 75 or applying a
hydraulic pressure at room temperature on the Br2/CH3CN-prepared fully hydrate phase. The
c ≈ 19.6 Å phase can be recovered by being exposed to humidity for a certain period of time
as the Br2/CH3CN-prepared fully hydrate phase transforms to the c ≈ 13.8 Å phase.
Nevertheless, the KMnO4-prepared c ≈ 13.9 Å phase of the high X samples cannot be
transformed to the c ≈ 19.6 Å phase. Formation of this new phase is ascribed to the sodium
site being replaced partially or mostly by the larger size of potassium, which will be discussed
in the later section.
In case of using aqueous NaMnO4 solution, the c ≈ 19.6 Å phase can be readily obtained
as well (see Fig. 5). Unlike the KMnO4 case, the high molar ratio of NaMnO4/Na would not
produce the c ≈ 13.9 Å phase.

Fe Kα

40x
Intensity (arb. units)

10x
(002)

(004)

5x
(006)

(100)
(102)
(103)
(104)

(106)

10 20 30 40 50 60
2θ (deg)

Figure 5. Powder x-ray diffraction patterns of Nax(H2O)yCoO2 prepared by immersing γ-


Na0.7CoO2 in aqueous NaMnO4 solution. The label 5X refers to that the molar ratio of NaMnO4 relative
to the Na in the parent compound is 5, i.e., 3.5 mole of NaMnO4 is used for1 mole of γ-Na0.7CoO2. All
the samples indicate the fully hydrated phase with c ≈ 19.6 Å.
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 11

3.4 FORMATION MECHANISM OF (Na,K)x(H2O)yCoO2


As previously stated, formation of superconductive Na0.35(H2O)1.3CoO2 is generally
considered as an oxidative de-intercalation process via the Br2/CH3CN route, followed by a
hydration process. For using the KMnO4 route, the direct formation of c ≈ 13.9 Å phase
involves not only the de-intercalation and hydration process but also an ion exchange
reaction. Furthermore, we find that the oxidative de-intercalation can be achieved simply by
immersed the parent compound in the water or hydrogen peroxide.

3.4.1 DE-INTERCALATION AND ION EXCHANGE

Table I summarizes the chemical compositions obtained by the ICP-AES analyses and
the lattice parameters of KMnO4-prepared (Na,K)x(H2O)yCoO2. Chemical analyses show that
the sodium content in (Na,K)x(H2O)yCoO2 decreases significantly at the molar ratio of
KMnO4/Na equal to or greater than 4.286, whereas the potassium content increases with
increasing molar ratio of KMnO4/Na. The potassium almost replaces the sodium completely
for the 40X sample. Besides, the sum of the sodium and potassium falls in the range of 0.28
and 0.38 in terms of the molar ratio with respect to cobalt. In an electrochemical study, [17] it
is found that the chemical potential of the sample increases with the increasing amount of Na
removed from the host lattice, meaning that to remove more Na from the host lattice requires
a higher redox potential. According to Nernst equation,

RT ⎛ aOx ⎞
E = E0 + ln⎜ ⎟⎟
zF ⎜⎝ a Re d ⎠, (1)

where E is the potential for a particular half-cell reaction, a is the activity of oxidized or
reduced species, and z is the number of electrons participating in the half-cell reaction.
Therefore, a higher redox potential would be expected for higher concentration of KMnO4
solution and consequently would result

Table I. Chemical analysesa of ICP-AES results and lattice constantsb of (Na,K)x(H2O)yCoO2


synthesized using aqueous KMnO4 solution.

Molar ratio
x =
of K Na Mnc Co a c
Na+K
KMnO4/Na (Å) (Å)
0.05X 0.00 0.38 0.38 0.04 1 2.8268(5) 19.626(2)
0.1X 0.01 0.34 0.35 0.07 1 2.8261(3) 19.573(1)
0.3X 0.02 0.33 0.35 0.08 1 2.8249(1) 19.669(1)
0.4X 0.02 0.34 0.36 0.08 1 2.8255(1) 19.686(1)
0.5X 0.02 0.34 0.36 0.08 1 2.8248(1) 19.679(1)
1X 0.03 0.35 0.38 0.08 1 2.8281(3) 19.685(2)
1.529X 0.05 0.31 0.36 0.07 1 2.8115(15) 19.750(5)
2.29X 0.08 0.28 0.36 0.07 1 2.8250(3) 19.735(2)
4.286X 0.18 0.15 0.33 0.07 1 2.8286(1) 13.917(1)
12 Chia-Jyi Liu

10X 0.21 0.07 0.28 0.08 1 2.8261(2) 13.864(1)


20X 0.26 0.05 0.31 0.08 1 2.8252(12) 13.849(1)
40X 0.25 0.03 0.28 0.08 1 2.8251(2) 13.960(1)
a
The error in weight % of each element in ICP-AES analysis is ±3 %, which corresponds to an
estimated error of ±0.02 per formula unit for each element.
b
The lattice constants are obtained using Rietica, a Rietveld structure refinement program, based on
a hexagonal structure (P63/mmc).
c
The manganese persistently exists in the sample after thoroughly washing procedure. The
manganese might exist as the amorphous form of MnO2 based on the reaction,
z → z
4MnO 4− + (2 + )H2O + ze−←4MnO 2 (s) + (3 − )O2(g) + (4 + z)OH−
2 4 ; [38] however, MnO2 can be partially
removed by immersing the sample in alkaline solution to form brown Mn(OH)3 precipitate.

in de-intercalating more alkaline metal from the host lattice, which is confirmed by the
smaller alkaline metal content (the sum of the sodium and potassium contents) in the higher X
samples.
Fig. 6 shows the XRD patterns of Na0.35CoO2 obtained by immersing γ-Na0.7CoO2 in the
30% H2O2 for 5 days, followed by storing a wet chamber for various periods of time. The
(002) reflection is found to shift to a lower angle at 2θ = 20.12° – 20.08° (d = 5.542-5.553 Å)
as compared to 2θ = 20.44° (d = 5.456 Å) for γ-Na0.7CoO2. The longer time the sample is
stored in the wet chamber, the lower 2θ the (002) reflection would occur. According to the
ICP-AES analysis, the sodium content is 0.35 which gives the evidence of de-intercalation
from the host lattice after the treatment of hydrogen peroxide. The water content is estimated
to be about 0.33 based on the TGA analysis shown in Fig. 7. As a result of de-intercalation of
Na and hydration, the c-axis expands slightly. In fact, simply by immersing γ-Na0.7CoO2 in
tap water, de-intercalation of Na can be achieved to have the composition of
Na0.30(H2O)yCoO2 and the (002) reflection occurs at 2θ = 20.02° (d = 5.569 Å) and the water
content of y ≈ 0.41 (Fig. 7).
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 13

Fe Kα
in wet chamber 44 days
Intensity ( arb.units)

in wet chamber 27 days

in wet chamber 7 days

in wet chamber 2 days

as -prepared

10 20 30 40 50 60
2 θ (degree)

Figure 6. Powder x-ray diffraction patterns of Nax(H2O)yCoO2 prepared by immersing γ-


Na0.7CoO2 in 30% H2O2 solution for different periods of time. The (002) reflection appears at 2θ =
20.12° – 20.08° indicating slight elongation of the c-axis as compared to 2θ = 20.44° for γ-Na0.7CoO2.
14 Chia-Jyi Liu

101
o
y=0.33 (37.4 C,100%)
100 o
y=0.28 (240 C,99.160%)
99

98
Weight% (%)

97
o
y=0.18 (280 C,97.471%)
96

95

94 o
y=0 (320 C,94.363%)
93
80 160 240 320 400 480 560
o
Temperature( C)

Figure 7. TGA curve of Na0.35(H2O)yCoO2 prepared by immersing γ-Na0.7CoO2 in 30% H2O2


solution for 5 days.
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 15

102
o
y=0.41 ( 31 C , 100%)
100 o
y=0.36 ( 50 C , 99.19%)
weight% (%)

98 o
y=0.32 ( 250 C , 98.45%)

96
o
y=0.19 ( 290 C , 96.23%)

94 o
y=0 ( 340 C , 92.99%)

92
50 100 150 200 250 300 350
o
Temperature ( C )

Figure 8. TGA curve of Na0.30(H2O)yCoO2 prepared by immersing γ-Na0.7CoO2 in tap water for 5
days.

3.4.2 HYDRATION

On contact with water vapor or with liquid water, AxMS2 (A = group 1A metal, M = Ti,
Nb, or Ta) undergoes a topotactic process and forms hydrated compounds Ax(H2O)yMS2 with
water molecules intercalated into the host lattice AxMS2. As a result, the c-axis expands with
little change in the a-axis. The size of the c-axis of is associated with the size of the alkaline
metal in the Ax(H2O)yMS2. [22] It has been found that the alkaline metal with the ionic radius
>1 Å leads to bilayers of water intercalated between the MS2 layers and hence a larger c-axis,
whereas the alkaline metal with the ionic radius <1 Å leads to monolayer of water inserted
within the alkaline metal layers and hence a smaller c-axis. The c-axis of A0.3(H2O)yTaS2 is
23.63 Å and 18.18 Å for A = Na+ (ionic radius: ~ 1 Å) and K+ (ionic radius: ~ 1.4 Å),
respectively. The interlayer height between MS2 layers is ca. 5.77 Å and 3.04 Å for A = Na+
and K+, respectively. They correspond well to once and twice the van der Waals diameter of a
water molecule of ~ 2.8 Å. Similar relationships between the size of the alkaline metal and
the c-axis could be found in the KMnO4-prepared cobalt oxyhydrates, as shown in Figs. 2 and
3 and Table I.
16 Chia-Jyi Liu

Due to the presence of potassium, the following scheme is proposed to describe the
formation of the new phase (Na,K)x(H2O)yCoO2 involving an ion exchange reaction in
addition to the de-intercalation and hydration process.

Scheme 1a

a b
Na0.7CoO2 Na0.7-zCoO2 (Na,K)xCoO2

(Na,K)x(H2O)yCoO2

a
(a) oxidative de-intercalation process: MnO4-(aq) at room temperature; (b) ion exchange reaction:
+
K (aq) at room temperature and x < 0.7; (c) hydration process: H2O at room temperature.

The whole process is completed as a one-pot reaction. Based on the XRD and ICP-AES
results, the c ≈ 13.9 Å phase having more K+ content has an interlayer expansion of ca. 1.5 Å,
whereas the c ≈ 19.6 Å phase having less K+ content has an interlayer expansion of ca. 4.35
Å. Direct formation of c ≈ 13.9 Å phase of cobalt oxyhydrate could be attributed to the
significant increase of K+ in the ≥ 4.286X samples, since a larger size of alkaline metal would
lead to a monolayer of water inserted within the alkaline metal layers, as discussed above in
the case of Ax(H2O)yMS2. It is therefore conceivable that formation of the two distinct c-axis
phases of cobalt oxyhydrates is associated with the size of the alkaline metal (K+ ~ 1.4 Å as
compared to Na+ ~ 1 Å). Further evidence can be seen in the TGA results, shown in Figs. 9
and 10. The c ≈ 13.9 Å phase has the hydration water content of y ≈ 0.7, while the c ≈ 19.6 Å
phase has y ≈ 1.33. It should be noted that the water content was found to be y ≈ 0.5 - 0.7 for
A = K and y ≈ 1.6 – 1.8 for A = Na in Ax(H2O)yMS2. In addition, the TGA curves show the
thermal stability of the water molecules in the lattice for the 0.3X sample is different from
that of the 10X sample. This could be due to the variation of the hydration energy for the Na+
and K+ and the position of the water molecules in the lattice. The water molecules are located
both within the Na layers and in between

3.5 KINETICS OF THE TOPOTACTIC TRANSFORMATION

The transformation from the non-hydrate phase (c ≈ 11.1 Å) to the fully hydrated phase
(c ≈ 19.6 Å) in both of the 0.05X and 0.1X samples suggests that the whole process is slow at
low concentration of KMnO4 solution. It is known that the oxidation reaction follows the
second order kinetics at low concentration of KMnO4 solution, which means the reaction rate
depends on both the concentration of the oxidized compound and the aqueous KMnO4
solution. At low concentration of KMnO4 solution, this could make the oxidative
deintercalation slow, and in turn slow down completing the formation of fully hydrated phase.
This is evidenced by the fact in the systematic study by immersing γ-Na0.7CoO2 in different
concentrations of aqueous KMnO4 solution for different periods of time. As shown in Fig. 11,
the c ≈ 19.6 Å phase has already appeared after one-day immersion for the 0.3X (the second
category) and 1X sample (the third category). The 0.05 X sample (the first category) appears
only as the non-hydrate phase (c ≈ 11.2 Å). Moreover, as shown in Fig. 12, the c ≈ 19.6 Å
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 17

phase shows up after 2 days of immersion for both the 0.05X sample. It can be readily seen
that the amount of non-hydrate phase existing in the sample decreases with increasing X
based on the relative reflection intensity of

100 . o
y=1.33 (26.3 C)

95 0.3X
o
. y=0.67 (44.6 C)
Wt% (%)

90
o
. y=0.34 (107.9 C)
85

80
. o
y=0 (250 C)

75
80 160 240 320 400 480 560
o
Temperature ( C)

Figure 9. TGA for the 0.3X sample obtained by immersing Na0.7CoO2 in the aqueous KMnO4
solution with the molar ratio of KMnO4/Na = 0.3.
18 Chia-Jyi Liu

100
o
y=0.73 (25.9 C)

10X
o
y=0.43 (62.7 C)
95
Wt %(%)

o
y=0.13 (144.5 C)

o
y=0 (250 C)
90

85
80 160 240 320 400 480 560
o
Temperature ( C)

Figure 10. TGA for the 10X sample Na0.07K0.21(H2O)yCoO2 obtained by immersing
Na0.7CoO2 in aqueous KMnO4 solution with the molar ratio of KMnO4/Na = 10. The water
content is determined by assuming the complete dehydration occurring at 250 .

Imax(c ≈ 19.6 Å)/Imax(c ≈ 11.2 Å). Besides, for the 0.05X sample already immersed in the
KMnO4 solution for 5 days, a small amount of non-hydrate phase still remains after additional
4 days of exposure to the humidity. All these results suggest that the hydration reaction in
intercalating water molecule into the host lattice is a slow process. For a higher X sample (≥
0.3X), the non-hydrate phase would not be observed after 5 days of KMnO4 treatment.
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 19

immersion for one day

(002)
1.0X

fully hydrate phase


Intensity (arb. units)

0.3X

(002)

non-hydrate phase

0.05X

5 10 15 20 25 30 35 40
2θ (degree)

Figure 11. Powder x-ray diffraction patterns of the 0.05X, 0.3X and 1X samples which are
obtained by immersing γ-Na0.7CoO2 in aqueous KMnO4 solution for 1 day. The fully hydrate phase (c ≈
19.6 Å) is observed both in the 0.3X and 1.0X sample, indicated by the reflection peak at 2θ ≈ 11.3°,
but not for the 0.05X sample. It suggests that formation of the fully hydrate phase is a very slow
process when using the low concentration of aqueous KMnO4 solution.
20 Chia-Jyi Liu

(002)
0.05X (immersion for 2 days)
Intensity (arb. units)

0.3X (immersion for 2 days)

fully-hydrate phase (002)

non-hydrate phase
0.05X (immersion for 3 days)

0.3X (immersion for 3 days)

5 10 15 20 25 30 35 40
2θ (degree)

Figure 12. Powder x-ray diffraction patterns of the 0.05X and 0.3X samples obtained by
immersing γ-Na0.7CoO2 in aqueous KMnO4 solution for 2 days and 3 days, respectively. The fully
hydrate phase (c ≈ 19.6 Å) appears after 2-day immersion in the aqueous KMnO4 solution for the 0.05X
sample. Formation of the c ≈ 19.6 Å phase for the 0.05X sample is much slower than that of the 0.3X
sample.

4. PHASE STABILITY
The fully hydrated phase is known to be very unstable in the ambient conditions. [21, 23]
This situation would obviously affect the interpretation of transport property measurements.
Even though with the nature of its instability, the c ≈ 19.6 Å phase of KMnO4-prepared
(Na,K)x(H2O)yCoO2 is found to be relatively more stable than those obtained by the
Br2/CH3CN solution. The c ≈ 13.9 Å phase of KMnO4-prepared (Na,K)x(H2O)yCoO2 is
relatively stable in the ambient conditions.

4.1 PHASE STABILITY OF THE c ≈ 19.6 Å PHASE OF (Na,K)x(H2O)yCoO2


Fig. 13 shows the XRD pattern evolution of the 0.3X sample. It can be seen that the XRD
peaks of the as-prepared wet powders, which are measured 40 min. after the washing and
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 21

filtration, are larger in intensity and sharper in peak width as compared to those of being
stored in the ambient air atmosphere for 24 h. This result indicates that the phase crystallinity
is deteriorating due to gradual loss of water molecule from the host lattice when storing the
sample in the ambient atmosphere. However, the fully hydrate phase can be brought back
when stored in a wet chamber with relative humidity of 98 % for 4 days (Fig. 13c). After
taken out from the wet chamber, the XRD peaks of the sample become smaller and broader in
a relatively short time shown in Figs. 13(d) and 13(e). Fig. 14 shows that the c ≈ 19.6 Å phase
obtained by the Br2/CH3CN solution readily deteriorates in a shorter period of time as
compared to that obtained by the aqueous KMnO4 solution.
The c ≈ 19.6 Å phase is also unstable when subjected to an evacuated environment.
Figs.15 and 16 compare the phase stability of the KMnO4-treated and Br2/CH3CN-treated
sodium cobalt oxyhydrates in different evacuated conditions. The Br2/CH3CN-treated c ≈
19.6 Å phase readily deteriorates in vacuum in a very short period of time. Knowing this fact
is important when characterizing the sample in a closed cycle refrigerator, which normally
requires evacuating the sample chamber. It seems that the KMnO4-treated c ≈ 19.6 Å phase
can sustain for a longer time in a similar vacuum condition and is more stable than the
Br2/CH3CN-treated one.

Fe Kα 0.3X
(a)
as-prepared wet powders
(002)

after filtration
(004)

(006)
Intensity (arb. uits)

(b) 24 hrs in air after filteration

(c)
1 day in wet chamber

(d) 30 min in air after taken out from wet chamber

(e) 2 hrs in air afte taken out from wet chamber

5 10 15 20 25 30 35 40
2θ (degree)

Figure 13. Evolution of XRD patterns for 0.3X sample obtained by immersing Na0.7CoO2 in
aqueous KMnO4 solution with the molar ratio of KMnO4/Na = 0.3. The XRD pattern is obtained for (a)
the as-prepared wet powders 35 min. after filtration; (b) the (a) powders stored in the ambient air for 24
h; (c) the (b) powders stored in a wet chamber with a relative humidity of 98 % for 1 day; (d) the (c)
powders taken out from wet chamber and stored in ambient air for 30 min; (e) the (c) powders taken out
from wet chamber and stored in ambient air for 2 h.
22 Chia-Jyi Liu

Fe Kα
Intensity (arb. units)

as-prepared powders

exposed in air for 40 mins

explosed in air for 11 hrs

5 10 15 20 25 30 35 40 45
2θ (degree)

Figure 14. Evolution of the XRD patterns for the Br2/CH3CN-prepared (Na,K)x(H2O)yCoO2. The c
≈ 19.6 Å phase readily deteriorates with the (002) reflection disappears in 11 hrs and a broad peak
emerges at 2θ ≈ 16°, where the (002) reflection of the c ≈ 13.9 Å phase occurs.
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 23

Intensity (arb. units)

as-prepared fully hydrated phase

mechanical pumping for 10 min

-5
6.6 x 10 torr for 7 mins

-5
1.9 x 10 torr for 30 min.

10 30 50 70 90
2θ (degree)

Figure 15. The evolution of XRD patterns of the KMnO4-treated 0.4X sample in different
evacuated conditions. The crystallinity of the c ≈ 19.6 Å phase gradually gets worse with higher
vacuum but still remains.
24 Chia-Jyi Liu

Intensity (arb. units)

fully hydrated

-4
1.6 x 10 torr for 4 min.

-5
9 x 10 torr for 7 min.

-5
2.7 x 10 torr for 2 hr

10 20 30 40 50 60 70 80 90
2θ (degree)

Figure 16. Evolution of the XRD patterns for the Br2/CH3CN-prepared (Na,K)x(H2O)yCoO2 in
different evacuated conditions. The c ≈ 19.6 Å phase readily deteriorates with evacuation. The (002)
reflection disappears in 11 hrs and a broad peak emerges at 2θ ≈ 16°, where the (002) reflection of the c
≈ 13.9 Å phase occurs.

As shown in Fig. 9, there is a multistage loss of weight as a result of loss of water upon
heating, indicating thermally unstable nature for the 0.3X sample. The 0.3X full hydrate
phase can be converted to the intermediate hydrate phase with c ≈ 13.9Å and the de-hydrate
phase with c ≈ 11.2 Å upon heating to 90 and 220 for a couple of hours, as shown in Fig.
17.

4.2 PHASE STABILITY OF THE c ≈ 13.9 Å PHASE OF (Na,K)x(H2O)yCoO2


Fig. 18 shows the evolution of XRD patterns for the 10X samples. The XRD patterns
remain unchanged after storing the samples in the wet chamber without any indication of
changing the phase to the c ≈ 19.6 Å phase. After taken out of the wet chamber and stored in
the ambient air for 2 days, the peak intensity has a slight decrease but remains unchanged
after exposed to the ambient air for a long period of time, suggesting it is a relatively stable
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 25

phase. This situation is remarkably different from the cases in the low X samples. As shown
in Fig. 19, the as-prepared wet powders contain the non-hydrate phase of c ≈ 11.2 Å and
convert to c ≈ 19.6 Å phase after storing the samples in the wet chamber for 4 days. The 10X
sample also shows a multistage loss of water (Fig. 10). The thermally unstable nature still
remains even with the Na mostly replaced by the K. As shown in Fig. 20, the XRD pattern
indicates that the pure de-hydrate phase cannot be obtained for the 4.286X sample heated at
the temperature as high as 600 . Above 300 , part of the sample has decomposed with the
appearance of Co3O4. Note that for the 0.3X sample and samples obtained by the Br2/CH3CN
solution, the de-hydrate phase can be readily obtained by heating at 220-250 . This result
indicates the potassium-containing cobalt oxyhydrates exhibit different thermal behavior from
the sodium cobalt oxyhydrates. This could be ascribed to the difference between the ion-
dipole interaction of K+-H2O within the alkaline layers and that of Na+-H2O located between
the CoO2 layers and the alkaline layers.

Fe Kα
fully hydrate
Intensity (arb. units)

intermediate hydrate

dehydrate

10 20 30 40 50 60
2θ (degree)

Figure 17. The XRD patterns of full hydrate (c ≈ 19.6 Å), intermediate hydrate (c ≈ 13.9 Å), and
dehydrate (c ≈ 11.2 Å) of Na0.33K0.22(H2O)yCoO2. The intermediate hydrate phase and dehydrate phase
were obtained by heating the fully hydrate phase at 90 and 220 , respectively.
26 Chia-Jyi Liu

(a) Fe Kα
as-prepared wet powders after filtration

(002)

(004)
Intensity (arb. units)

(b) 14 hrs in air after filtration

(c) 4 days in in wet chamber

(d) 2 days in air after taken out from wet chamber

5 10 15 20 25 30 35 40
2θ (degree)

Figure 18. Evolution of XRD patterns for 10X sample obtained by immersing Na0.7CoO2 in
aqueous KMnO4 solution with the molar ratio of KMnO4/Na = 10. The XRD pattern is obtained for (a)
the as-prepared wet powders 40 min after filtration; (b) the (a) powders stored in the ambient air for 14
h; (c) the (b) powders stored in a wet chamber with relative humidity of 98 % for 4 days; (d) the (c)
powders stored in the ambient air for 2 days after taken out from the wet chamber.
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 27

Fe Kα 0.05X
*

(a) *
as-prepared wet powders
Intensity (arb. units)

after filtration

(b) 24 hrs in air after filtration

(c)
(002)

(004)
4 days in wet chamber

(006)
*

5 10 15 20 25 30 35 40
2θ (degree)

Figure 19. Evolution of XRD patterns for 0.05X sample obtained by immersing γ-Na0.7CoO2 in the
aqueous KMnO4 solution with the molar ratio of KMnO4/Na = 0.05. The XRD pattern is obtained for
(a) the as-prepared wet powders 40 min after filtration; (b) the (a) powders stored in the ambient air for
24 h; (c) the (b) powders stored in a wet chamber with relative humidity of 98 % for 4 days. The
asterisk indicate the (002) reflection of the de-hydrated c ≈ 11.2 Å phase.
28 Chia-Jyi Liu

* Co3O4 * o
750 C 2 hrs in N
2
* *
* * *
Intensity(arb. units)

o
* 500 C 6 hrs in air
* * * * *

o
* 400 C 6 hrs in N
2
* * * * *

o
* 300 C 12 hrs in air
*
* * *

10 20 30 40 50 60 70 80 90
2θ (degree)

Figure 20. The XRD patterns for the 4.286X sample heated at various temperatures with attempt
to obtain the de-hydrate phase. However, the pure de-hydrate phase cannot be obtained at different
combinations of temperature and atmosphere. The c ≈ 13.9Å phase persistently coexists with the
dehydrate phase and the sample has decomposed partially accompanied by the formation of Co3O4,
indicating different thermal stability from the low X samples.

5. PHYSICAL CHARACTERIZATION OF (NA,K)X(H2O)YCOO2

5.1 SUPERCONDUCTIVITY
Long period of exposure to the water vapor for the 0.1X sample results in the growth of
the c ≈ 19.6 Å phase at the expense of the non-hydrate phase (c ≈ 11.2 Å). This phase
transformation also changes the onset superconducting transition temperature of the 0.1X
sample from 3.4 K to 4.5 K, shown in Fig. 21. Apparently, a larger portion of c ≈ 19.6 Å
phase in the sample would result in a higher onset superconducting transition temperature.
This could be caused by superimposition of the diamagnetic signal of a superconductor (c ≈
19.6 Å phase) and the paramagnetic signal of a non-superconductor (c ≈ 11.2 Å). The positive
paramagnetic signal would shift the onset Tc to a lower temperature when the c ≈ 11.2 Å
phase is present in the sample. It should be noted that the sodium content of the 0.1X sample
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 29

is not expected to change after being exposed to the water vapor. This might pose a question
on the dependence of superconducting transition temperature on the sodium content. [24] Fig.
22 shows the zero-field cooled magnetization data for the superconducting KMnO4-treated
samples containing the 19.6 Å phase. The onset superconducting transition, defined as the
magnetization starting to decrease, is 3.2 K, 3.4 K, 4.6 K, 4.6.K, 4.5 K, and 3.5 K for the
0.05X, 0.1X, 0.3X, 0.5X, 1.529X, and 2.29X samples, respectively. The 2.29X sample shows
an increasing magnetization with decreasing temperature before undergoing superconductive
transition, which could be ascribed to the existence of the c ≈ 13.9 Å phase. Fig. 23 shows the
zero-field cooled mass magnetization data obtained in an applied field of 20 Oe for the
KMnO4-treated samples having the pure 13.9 Å phase. It can be readily seen that the 13.9 Å
phase of (Na,K)x(H2O)yCoO2 would not undergo superconductive transition down to 2 K.

5.2 THERMOPOWER
5.2.1 INTRODUCTION
Thermopower (TEP) measurements can provide both information of the type and the
characteristic energy of charge carriers, which elucidate the conduction process and
thermodynamics. Since TEP is a measure of the heat per carrier over temperature, we can
thus view it as a measure of the entropy per carrier. For NaxCoO2, the large thermopower has
been ascribed to the spin and orbital degeneracy due to the competition of crystal field and
Hund’s rule coupling in cobalt ions. [25,26] Heat capacity, [27] magnetic, [28] thermopower,
[29] angle-resolved photoemission spectroscopy (ARPES) [30], and

0.005
as-prepared 0.1X
0.1X ( 10 days in a wet chamber)

0
M (emu/g)

Fe Kα
Intensity (arb. units)

-0.005 as-prepared 0.1X


*

0.3X 10 days in a wet chamber


-0.01
(002)

(004)

(006)

5 10 15 20 25 30 35 40
2 θ (degree)
-0.015
0 2 4 6 8 10 12 14 16
Temperature (K)

Figure 21. The zero-field cooled dc magnetization data of the as-prepared 0.1X sample, after-
exposed-to-humidity 0.1X and the 0.3X samples. The applied field is 10 Oe and 20 Oe for 0.3X and
0.1X, respectively. The inset shows the growth of the c ≈ 19.6 Å phase of the 0.1X sample at the
expense of the non-hydrate phase after 10-day exposure to the humidity, as evidenced by the
disappearance of the (002) peak (*) of the non-hydrate phase (c ≈ 11.2 Å).
30 Chia-Jyi Liu

0.01
2.29X
Magnetization (emu/g)

0.005
0.05X
0
0.1X

-0.005
1.529X

-0.01 0.5X

0.3X
-0.015
0 5 10 15 20
Temperature (K)

Figure 22. Zero-field cooled mass magnetization data for the low X samples containing the 19.6Å
phase. The 0.05X and 0.1X are measured in a field of 20 Oe and the rest of the samples are measured in
a field of 10 Oe. Note that only 0.3X sample is the pure 19.6Å phase. The 0.5X, 1.529X, and 2.29X
sample are a mixture of 19.6Å and 13.9Å phase and the 0.05X and 0.1X samples are a mixture of
19.6Å and 11.2Å phase.
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 31

0.0025

0.002 40X

20X
M (emu/g)

0.0015

0.001

0.0005
10X
0
0 10 20 30 40 50 60
Temperature (K)

Figure 23. Zero-field cooled mass magnetization data obtained in an applied field of 20 Oe for the
high X samples containing the pure 13.9Å phase.

polarization-dependent soft x-ray absorption [31] measurements have suggested a strong


correlation of 3d electrons in NaxCoO2. In a strong correlated system based on the Hubbard
model, the thermopower can be expressed as

1 S ( 2) µ
S =− (1)
+
T S eT , (2)

where e is the absolute value of the electron charge and

1 e ∞ ⎧⎪ ⎡ 1 ⎛ ⎞⎤ ⎫⎪
S(2) = ∫0 ⎨Tr⎢exp ⎜ µ∑ ni − H ⎟⎥ ×[Qv(τ ) + v(τ )Q]⎬dτ
2 kBT ⎪⎩ ⎣ kBT ⎝ i ⎠⎦ ⎪⎭
,

1 e ∞ ⎧⎪ ⎡
2
1 ⎛ ⎞⎤ ⎫⎪
S (1) = ∫0 ⎨Tr⎢exp ⎜ µ∑ ni − H ⎟⎥ × [vv(τ ) + v(τ )v]⎬dτ
2 kBT ⎪⎩ ⎣ kBT ⎝ i ⎠⎦ ⎪⎭
,
(3)
32 Chia-Jyi Liu

where “Tr” denote the trace and is the summation over some complete set of states. H, µ, ν
and Q are the Hamiltonian, chemical potential, velocity and energy flux operators,
respectively. In the high temperature limit, thermopower is dominated by the entropy
µ S ( 2)
(1)
term eT in Eq. (2) since the energy-transport term TS is small under the condition of
U
→∞
strong Coulomb interaction with T . [32,33] For the transition metal oxides, the room
temperature thermopower can be considered close to the high temperature limit. [34, 35]

5.2.2 THERMOPOWER OF (Na,K)x(H2O)yCoO2

As shown in Fig. 17, Na0.33K0.02(H2O)yCoO2 can have three phases with the same metal
composition and hexagonal structure except the variation in the c axis and the water content,
i.e., the full hydrate phase (y ≈ 1.33) with c ≈ 19.6 Å, the intermediate hydrate phase (y ≈
0.67) with c ≈ 13.9 Å and the dehydrate phase (y = 0) with c ≈ 11.2 Å. Since the alkali metal
(Na,K) layer is considered as a charge reservoir, the average valence of Co is not expected to
change by the variation of water content. Larger separation of trigonal CoO2 layer is
expected to enhance two-dimensional (2D) character. One should expect their physical
properties will be affected by the anisotropy with respect to the bonding and structure.
As mentioned previously, the c ≈ 19.6 Å phase is unstable in the ambient environment.
One has to take precaution when handling the sample for physical characterization. In
particular, transport property measurements normally require pressing the sample into pellets.
The c ≈ 19.6 Å phase could easily transform into the c ≈ 13.9 Å phase when applying too
high a pressure on the sample. Besides, as shown in Fig. 24 the c ≈ 19.6 Å phase could
transform into the c ≈ 13.9 Å phase during the lengthy measuring process. As shown in Fig.
25, a device is designed to overcome this problem by keeping the measuring chamber with
sufficient humidity and preventing the measured sample from losing water. Fig. 26 shows that
the c ≈ 19.6 Å phase remains as it is after lengthy measurements.
In a strong correlated system with no double occupancy, namely, t « kBT « U, where t is
the transfer integral of an electron between neighboring sites and U is on-site Coulomb
interaction, the thermopower in the high-temperature limit can be expressed by the
generalized Heikes formula as [32]
kB ⎛ 1− ρ ⎞
S =− ln⎜2 ⎟
e ⎜⎝ ρ ⎟⎠
, (4)

where ρ = n/N is the ratio of particles to sites (n: the number of particles; N: the number of
available sites). For the strong correlated electron system of NaxCoO2, considering the effects
of spin and orbital
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 33

Intensity (arb. units)

before measurement

after measurement

10 20 30 40 50 60
2θ (degree)

Figure 24. The c ≈ 19.6 Å phase transforms into the c ≈ 13.9 Å phase after thermopower
measurements.

Figure 25. The thermopower measuring device maintains sufficient humidity by sealing up a pan
of water in the sample holder.
34 Chia-Jyi Liu

(002)

(004)
Intensity(arb. units)

Before measurement

(006)
After measurement

5 10 15 20 25 30 35 40
2θ(degree)

Figure 26. The c ≈ 19.6 Å phase still remains as it is after thermopower measurements.

degeneracy, the thermopower can be expressed as [34,35]

k B ⎛ g3 x ⎞
S=− ln⎜ ⎟
e ⎜⎝ g 4 1 − x ⎟⎠
, (5)

where x is the concentration of Co4+, g3 and g4 are the number of the configuration of Co3+
and Co4+ ions, respectively. The term of kB/e is 86.3 µV/K. Neglecting the possible non-
stoichiometry of the oxygen, the formal valence of the cobalt ion for the γ-Na0.7CoO2 and
Na0.33K0.02(H2O)1.33CoO2 is +3.3 and +3.65, respectively, based on a simple calculation of
charge neutrality. Co3.3+ corresponds to a mixed valence system comprising 70% of Co3+ and
30% Co4+ (x = 0.3); Co3.65+ corresponds to a mixed valence system comprising 35% of Co3+
and 65% Co4+ (x = 0.65). The spin state of cobalt ion could be low spin (LS), intermediate
spin (IS), or high spin (HS), being associated with the electronic state of cobalt ion. The
number of the configuration of cobalt ion is determined by the competition between the
crystal field splitting 10 Dq and the Hund's rule coupling as well as temperature. Table II lists
the possible spin state of Co3+ and Co4+, the ratio of g3/ g4, and the calculated thermopower
using Eq. (5) for γ-Na0.7CoO2 and Na0.33K0.02(H2O)1.33CoO2. Fig. 27 shows the thermopower
as a function of temperature for γ-Na0.7CoO2 and Na0.33K0.02(H2O)1.33CoO2. When compared
to the value of thermopower in Table II, both Co3+ and Co4+ of γ-Na0.7CoO2 and
Na0.33K0.02(H2O)1.33CoO2 seem to be a mixed spin system of coexisting HS, LS and IS with
their states close in energy. However, based on magnetization measurements, both Co3+ and
Co4+ of γ-Na0.7CoO2 should be in the LS state since the effective moment per cobalt in γ-
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 35

Na0.7CoO2 is estimated to be ca. 1.2 µB and the theoretical effective moment for Co3+ and
Co4+ in the LS state is 0 and 1.73 µB. [36] In both cases, the value of thermopower turns
smaller as the Co4+ concentration increases, which is qualitatively consistent with the data in
Fig. 27. Realizing that the valence state of cobalt ion for the c ≈ 13.9 Å phase is not different
from that for the c ≈ 13.9 Å phase, [37] the smaller value, shown in Fig. 28, of the
thermopower for the c ≈ 13.9 Å phase as compared to the c ≈ 19.6 Å phase should not be
ascribed to the change of valence state of cobalt. Instead, it should be associated with the
possible change of the electronic structure since thermopower is very sensitive to the
topology of the Fermi surface. The nonlinear behavior of the thermopower in both the c ≈
19.6 Å and the c ≈ 13.9 Å phases is likely due to the enhancement of electron-electron
correlation. If it is the case, the enhanced two-dimensionality in the c ≈ 19.6 Å phase could
play a role in affecting the electron-electron correlation and hence the extent of enhancement
of thermopower. The thermopower of the c ≈ 11.2 Å phase shows a distinct behavior from the
other two,

Table II. The possible spin state of Co3+ and Co4+, the ratio of g3/ g4, and the calculated
thermopower using Eq. (5) for γ-Na0.7CoO2 (x = 0.3)a and Na0.33K0.02(H2O)1.33CoO2 (x =0.65).

spin state of Co3+ spin state of Co4+ g3/g4 S(µV/K) x=0.3 S(µV/K) x=0.65
HS HS 15/6 -6 -132
HS+LS HS 16/6 -12 -138
HS HS+LS 1/12 288 161
LS LS 1/6 228 101
HS+LS+IS HS+LS+IS 34/36 78 48
a
The value of x is the concentration of Co4+.

100

γ-Na CoO
80 0.7 2

60
S (µV/K)

full hydrate 0.3X


40

20

0
0 50 100 150 200 250 300
Temperature (K)

Figure 27. The temperature-dependent thermopower of γ-Na0.7CoO2 obtained and


Na0.33K0.02(H2O)1.33CoO2 (0.3X) with c ≈ 19.6 Å. The γ-Na0.7CoO2 is prepared by the rapid-heat up
procedure at 700°C, followed by sintering at 800°C in air.
36 Chia-Jyi Liu

50
40
full hydrate
30
S (µV/K)

20
10 de-hydrate
intermediate
0 hydrate

-10
-20
0 50 100 150 200 250 300
Temperature (K)

Figure 28. The temperature-dependent thermopower of three Na0.33K0.02(H2O)yCoO2 phases with c


≈ 19.6 Å (full hydrate), c ≈ 13.9 Å (intermediate hydrate), and c ≈ 11.2 Å (de-hydrate), respectively.

40

35

30

25
S (µV/K)

20

15

10

0
0 50 100 150 200 250 300
Temperature (K)

Figure 29. The temperature-dependent thermopower of three Na0.07K0.21(H2O)0.73CoO2 phase with


c ≈ 13.864(1) Å.
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 37

indicating a dramatic change of the electronic structure. In the wet-chemical analyses, the c ≈
11.2 Å phase shows a decrease of the valence state of cobalt ion as compared to the other two
and has oxygen deficiency in the lattice caused by heating to obtain this phase. [37] In
particular, the sign change of thermopower suggests co-existence of electrons and holes,
which could be associated with the oxygen deficiency. Fig. 29 shows the temperature-
dependent thermopower of the KMnO4-treated 10X sample of Na0.07K0.21(H2O)0.73CoO2. Both
the size and the temperature dependence of the thermopower bear a resemblance to the c ≈
13.9 Å phase of Na0.33K0.02(H2O)yCoO2. Apparently, the potassium content would not affect
the thermopower behavior; instead, the key factors rely on the concentration of Co4+ and the
length of the c-axis (the dimensionality of the crystal structure). The Na0.07K0.21(H2O)0.73CoO2
phase with c ≈ c = 13.864(1) Å

CONCLUSION
Potassium sodium cobalt oxyhydrates (Na,K)x(H2O)yCoO2 can be synthesized using the
aqueous KMnO4 solution instead of using the toxic Br2/CH3CN route. By immersing the γ-
Na0.7CoO2 in aqueous KMnO4 solution with 0.05 ≤ KMnO4/Na ≤ 2.29, the superconductive
(Na,K)x(H2O)yCoO2 phase with c ≈ 19.6Å can be obtained. For KMnO4/Na ≥4.286, the c ≈
13.9Å phase of (Na,K)x(H2O)yCoO2 do not undergo superconductive transition down to 2 K.
Formation of the (Na,K)x(H2O)yCoO2 involves not only the oxidative deintercalation and
hydration but also the ion exchange reaction. High molar ratio of KMnO4/Na (≥4.286) leads a
pure c ≈ 13.9Å phase with a monolayer of water within the alkaline layer as a result of partial
or almost complete substitution of K+ for Na+. The hydration is a very slow process
particularly when using the aqueous KMnO4 solution with a low molar ratio of KMnO4/Na.
By using the aqueous NaMnO4 solution, the potassium-free superconductive Nax(H2O)yCoO2
phase can be also obtained. By simply immersing γ-Na0.7CoO2 in tap water or H2O2, de-
intercalation of sodium would occur with the sodium content of 0.3-0.4, which is about the
same as using the KMnO4 or the Br2 routes. As a result, it only leads to a slight expansion of
the c-axis (ca. 11.2Å), which is very close to that of the de-hydrate phase. However, there is a
sign for the appearance of the c ≈ 19.6Å phase by simply storing the γ-NaxCoO2 in a wet
chamber in our recent results.
The c ≈ 19.6Å phase of (Na,K)x(H2O)yCoO2 is unstable in the ambient air, evacuated
chamber, and the hydraulic pressure and can transform into the c ≈ 13.9Å phase. The
KMnO4-treated (Na,K)x(H2O)yCoO2 seems to be more robust than the Br2-treated one. The c
≈ 13.9Å phase of (Na,K)x(H2O)yCoO2 is quite stable when subjected to the above conditions.
The c ≈ 19.6Å phase of (Na,K)x(H2O)yCoO2 is also unstable and would lose water upon
heating accompanied by converting to the c ≈ 13.9Å phase and c ≈ 11.2Å phase at 90°C and
220°C, respectively.
For transport property measurements, one should check the XRD after measurements to
make sure the phase being the same as before measurements. The effects of the spin and
orbital degeneracy seem to play a role in the cobalt oxide system with strong electron-
electron correlation. The size of thermopower at room temperature for
Na0.33K0.02(H2O)1.33CoO2 decreases as compared to that for γ-Na0.7CoO2 due to the
increasing concentration of Co4+, which is qualitatively consistent with the calculations based
on the generalized Heikes formula. In spite of the fact that both the c ≈ 19.6Å and c ≈ 13.9Å
38 Chia-Jyi Liu

phases have the same concentration of Co4+, the variation of the thermopower could be
arising from the enhanced two-dimensionality of the c ≈ 19.6Å phase, which affects the
thermopower enhancement of the electron-electron correlation. The Na0.07K0.21(H2O)0.73CoO2
phase with c ≈ 13.864(1) Å exhibits similar thermopower behavior to the c ≈ 13.9Å phase of
Na0.33K0.02(H2O)yCoO2, suggesting that the content of potassium plays insignificant role in
the transport property and that the length of the c-axis, i.e. the dimensionality of the crystal
structure, has a role on the parameters such as electron-electron correlation, Hund's rule
coupling or energy level splitting between orbitals, affecting the transport properties.

Acknowledgment
This work is supported by the National Science Council of ROC, grant Nos. NSC 92-
2112-M-018-005 and NSC 93-2112-M-018-003. I would like to thank C.-Y. Liao, W.-C.
Hung, J.-S. Wang for the experimental works and acknowledge fruitful collaboration with Y.-
Y. Chen, S. Neeleshwar, H.-S. Sheu, and J.-Y. Lin.

REFERENCES
[1] Takada, K,; Sakurai, H.; Takayama-Muromachi, E.; Izumi, F.; Dilanian, R. A.;
Sasaki, T. Nature 2003, 422, 53.
[2] Terasaki, I.; Sasago, Y.; Uchinokura, K. Phys. Rev. B 1997, 56, R12685.
[3] Ando, Y.; Miyamoto N.; Segawa, K.; Kawata, T.; Terasaki, I. Phys. Rev. B 1999,
60, 10580.
[4] Ray, R.; Ghoshray, A.; Ghoshray, K.; Nakamura, S. Phys. Rev. B 1999, 59, 9454.
[5] Wang, Y.; Rogado, N. S.; Cava, R. J.; Ong, N. P. Nature 1999, 423, 425.
[6] Hasan, M. Z.; Chuang, Y.-D.; Kuprin, A. P.; Kong, Y.; Qian, D.; Li, Y. W.;
Mesler, B. L.; Hussain, Z.; Fedorov, A. V.; Kimmerling, R.; Rotenberg, E.;
Rossnagel, K.; Koh, H.; Rogado, N. S.; Foo, M. L.; Cava, R. J. (2003). Fermi
surface and quasiparticle dynamics of NaxCoO2 (x=0.7) investigated by angle-
resolved photoemission spectroscopy. cond-mat/0308438.
[7] Fouassier, C.; Matejka, G.; Reau, J.-M.; Hagenmuller, P. J. Solid State Chem.
1973, 6, 532.
[8] Wycoff, Ralph W. G. Crystal Structures; Inorganic Compounds RXn, RnMX2,
RnMX3; Robert E. Krieger Publishing Company: Malabar, FL, 1986; Vol. 2, pp
291-296.
[9] Parant, J.-P.; Glazcuaga, R.; Devalette, M.; Fouassier, C.; Hagenmuller, P. J. Solid
State Chem. 1971, 3, 1.
[10] Cushing, B. L.; Wiley, J. B. J. Solid State Chem. 1998, 141, 385.
[11] Ono, Y.; Ishikawa, R.; Miyazaki, Y.; Ishii, Y.; Morii, Y.; Kajotani, T. J. Solid State
Chem. 2002, 166, 177.
[12] Hahn, T. International Tables for Crystallography; Space-Group Symmetry; D.
Reidel Publishing Company: Dordrecht, Holland, 1983; Vol. A, pp 590-591.
Preparation and Thermopower Behavior of the New Cobalt Oxyhydrates 39

[13] Zhang, P.; Luo, W.; Crespi, V. H.; Cohen, M. L.; Loui, S. G. Phys. Rev. B 2004,
70, 085108.
[14] Zhang P.; Capaz, R. B.; Cohen, M. L.; Louie, S. G. (2005) Theory of Sodium
Ordering in NaxCoO2. cond-mat/0502072.
[15] Zandbergen, H. W.; Foo, M.; Xu, Q.; Kumar, V.; Cava, R. J. Phys. Rev. B 2004,
70, 085108.
[16] Liu, C.-J.; Liao, C.-Y.; Huang, L.-C.; Su, C.-H.; Neeleshwar, S.; Chen, Y.-Y.; Liu,
C.-J. C. Physica C 2004, 416, 43.
[17] Chou, F.C.; Cho, J. H.; Lee, P. A.; Abel, E. T.; Matan, K.; Lee Y. S. Phys. Rev.
Lett. 2004, 92, 157004.
[18] Kawata, T.; Iguchi, Y.; Ito, T.; Takahata, K.; Terasaki, I. Phys. Rev. B 1999, 60,
10584.
[19] Liu, C.-J.; Liao, J.-Y.; Wu, T.-W.; Jen, B.-Y. J. Mater. Sci. 2004, 39, 4569.
[20] Motohashi, T; Naujalis, E.; Ueda, R.; Isawa, K; Karppinen, M.; Yamauchi, H.
Appl. Phys. Lett. 2001, 79, 1480.
[21] Foo, M. L.; Schaak, R. E.; Miller, V. L.; Klimczuk, T.; Rogado, N. S.; Wang, Y.;
Lau, G. C.; Craley, C.; Zandbergen, H. W.; Png, N. P.; Cava, R. J. Solid State
Commun. 2003, 127, 33.
[22] Lerf, A and Schöllhorn, R. Inorg. Chem. 1977, 16, 2950.
[23] Chen, D. P.; Chen, H. C.; Maljuk, A.; Kulakov, A.; Zhang, H.; Lemmens, P.; Lin,
C. T. Phys. Rev. B 2004, 70, 024506.
[24] Schaak, R. E.; Klimczuk, T.; Foo, M. L.; Cava, R. J. Nature 2003, 424, 527.
[25] Koshibae, W.; Tsutsui, K.; Maekawa, S. Phys. Rev. B 2000, 62, 6869.
[26] Koshibae, W.; Maekawa, S. Phys. Rev. Lett. 2001, 87, 236603.
[27] Ando, Y.; Miyamoto, N.; Segawa, K.; Kawata, T.; Terasaki, I. Phys. Rev. B 1999,
60, 10580.
[28] Ray, R.; Ghoshray, A.; Ghoshray, K.; Nakamura, S. Phys. Rev. B 1999, 59, 9454.
[29] Wang, Y.; Rogado, N. S.; Cava, R. J.; Ong, N. P. Nature 1999, 423, 425.
[30] Hasan, M. Z. Chunag, Y.-D.; Qian, D.; Li, Y. W.; Kong, Y.; Kuprin, A.; Fedorov,
A. V.; Kimmerling, R.; Rotenberg, E.; Rossnagel, K.; Hussain, Z.; Koh, H.;
Rogado, N. S.; Foo, M. L.; Cava, R. J. Phys. Rev. Lett. 2004, 92, 246402.
[31] Wu, W. B.; Huang, D. J.; Okamoto, J.; Tanaka, A.; Lin, H.-J.; Chou, F. C.;
Fujimori, A.; Chen, C. T. Phys. Rev. Lett. 2005, 94, 146402.
[32] Chaikin, P. M.; Beni, G. 1976, 13, 647.
[33] Beni, G. Phys. Rev. B 1974, 10, 2186.
[34] Oguri A.; Maekawa, S. Phys. Rev. B 1990, 41, 6977.
[35] Pálsson G.; Kotliar, G. Phys. Rev. Lett. 1998, 80, 4775.
[36] Wu, T.-W. (2003) Studies on the structure and thermoelectric properties of bronze
type cobalt oxides. Master thesis of National Changhua University of Education,
Taiwan.
[37] Karppinen, M.; Asako, I.; Motohashi, T.; Yamauchi, H. Chem. Mater. 2004, 16,
1693.
[38] Skoog, D. A.; West, D. M., Fundamental of Analytical Chemistry, 3rd ed. Holt,
Rinehart and Winston, Inc.: New York, 1976; p. 341.

Vous aimerez peut-être aussi