Вы находитесь на странице: 1из 6

Appl. Phys.

A 59, 173-178 (1994)

Applied

Ph~ic~ A and Surfaces


Springer-Verlag 1994

Solids

Lithium intercalation in MoO3: A comparison between crystalline and disordered phases


C. Julien, A. Khelfa, J. P. Guesdon, A. Gorenstein*
Laboratoire de Physique des Solides, associ~ au CNRS, Universit~ Pierre et Marie Curie, 4, place Jussieu, F-75252 Paris Cedex 05, France (Fax: + 33-1/44274541, E-mail: MI@frunip.62) Received 18 December 1993/Accepted21 April 1994

Abstract. We report properties of lithium-intercalated MoO3 crystalline and thin-film which are potential cathode materials for high energy density batteries. Discharge and charge reactions of MoO3 electrodes in a non-aqueous Li+-electrolyte have been studied. The kinetically accessible discharge range amounts to 0_<x_< 1.5 for Li insertion in LixMoO3. Transport parameters such as the Li + chemical diffusion coefficient, thermodynamic factor and ionic conductivity are investigated during the Li + insertion process and discussed with respect to the crystallinity of the cathode material. PACS" 68.65.+g, 66.30.-h

A number of transition-metal compounds based on either two-dimensional van der Waals bonded structures or three-dimensional framework tunnel structures are of interest as potential electroactive materials for high energy density secondary lithium batteries [1]. There have been numerous studies made on the layered dichalcogenides [2-5], layered chalcogenophosphates [6, 7], ternary sulphides [8, 9], oxyhalogenides [10], and oxides of transition metals [11, 12]. These materials undergo reversible topotactic insertion reactions with lithium. Among various transition-metal oxides, only a few have corrugated layered structures suitable for storagebattery electrodes [13]. Molybdenum trioxide, MOO3, forms an ideal host structure for the reversible insertion of Li ions and is a potential cathode material [13-20]. Owing to the high degree of utilization of MoO3 cathodes the practical energy density of the Li/MoOa couple exceeds that of the Li/TiS2 system [21]. Another advantage of MoO 3 is the possibility to prepare thin-films very easily. It has been pointed out by several workers that amorphous cathode active materials are very useful for im* Permanent address: Physics Department, Universidade Estadual de Campinas, CP 6165, Campinas, SP, Brazil

provement of rechargeability [22-27]. A typical example developed by Sakurai and Yamaki [26], and worth noting because of its simple method of preparation and cathode charge-discharge characteristics, is a V2Os-P20 5 glassy material, but a low rate of the ionic transport has been observed [28]. Thin-films are appropriate materials for studying the relation between Li + ion transport and the crystallinity of the substance. Thin films of MoO3 have been the subject of numerous investigations [29-31], but studies of lithium intercalation into MoO3 thin films are sparse [32, 33]. This paper presents investigations on the Li + ionic transport in both M o O 3 crystal and thin film. We have studied their electrochemical ability of lithium intercalation and determined chemical diffusion coefficients, thermodynamic factors and ionic conductivities for both forms of MOO3.

1 Experimental
Molybdenum trioxide possesses an intriguing layered structure consisting of double layers of MoO6 octahedra held together by covalent forces in the (100) and (001) directions but by much weaker van der Waals forces in the (010) direction as shown in Fig. 1. The important consequence of this is that large MoO3 crystallites have elongated slab-like habits, preferentially exposing the (010) direction corresponding to the b-axis in the orthorhombic symmetry.

1.1 Crystal preparation


Crystals of MoO3 were prepared by heating molybdic acid, MoO3-H20 (Prolabo). The powdered material was loaded in a platinum crucible and positioned in a twozone tube furnace. A purified argon flow of 50 cm3/min was maintained during the experiment. The powder was heat treated at 7 5 0 + 2 C for 10h; crystals were then

174

C. Julien et al.

ing at t = 0 obeyes the following equation E* = (IWRT/F2 c* A ) [(4tfizD ) I/2- t/ O], (1)

~..:.:" *" i';" " ~

where E* is the voltage of the cell with reference to the open circuit voltage, I is the current intensity, t is the time, c* is the Li + ion concentration in the active material, W is the thermodynamic factor, A and 6 are the surface and the thickness of the electrode, T is the temperature, F = 96 500 As mol- 1 and R = 8.314 J K - 1 mol - 1. The thermodynamic factor is defined as W = ~(ln a*)/ Q(ln c*) in which a* is the activity of the intercalant species in the solid solution electrode. Wis also expressed by the ratio
w = D/D0, (2)

between the chemical diffusion coefficient D, and the component diffusion coefficient Do.
Fig. 1, Schematic representation of the crystal structure of MoO3 showing infinite of MoO6 octahedra perpendicular to the b-axis

2 Results and discussion

formed in the second zone at about 650 C. Small crystals of M oOa have the shape of elongated platelets with (010) perpendicular to the basal ac-plane and (001) along the longest edge. 1.2 Film preparation Films of MoO3 were prepared by flash evaporation of pre-baked molybdenum oxide powder (Alpha Products, Puratronic 99.9995%) on silica glass substrates maintained in the temperature range 30-300 C in a vacuum better than 5 x 10 -6 Tort. During evaporation a boat of molybdenum was maintained at 1000 C and the rate of deposition was 4 nm/s. The thickness of the films was in the range 0.5-0.6 gm. Flash-evaporated MoOa films are uniformly thick with a blue colour. The colour of the samples suggests that the films contain a number of oxygen vacancies [34]. The colour of these samples varied from light blue to deep blue with increasing substrate temperature (Ts). This is attributed to the increase of the lack of oxygen atoms with increasing T~ during the deposition. 1.3 Electrochemical measurements Electrochemical measurements of lithium intercalation in MoO3 samples have been carried out by coulometric titration in Li/MoO3 galvanic cells with 1 M LiC104 in Propylene Carbonate (PC) solution as a non-aqueous electrolyte. The kinetic parameters, i.e., the chemical diffusion coefficient and the thermodynamic factor have been determined by the Modified Galvanostatic Intermittent Titration Technique (MGITT) from the variation of the cell voltage versus time during the relaxation period following a long discharge or charge [35]. This method states that the evolution of the voltage after circuit open-

MoO3 cathode materials have been characterized by X-ray diffraction measurements using a Cu K~ radiation. Figure 2 shows the X-ray diffraction patterns of MoO3 crystal and MoOa thin films. X-ray diffraction data of the MoO3 crystal indicate that molybdenum trioxide crystallizes in the orthorhombic structure with a = 0.3962 nm, b = 1.3858 nm and c=0.3696 nm. The diagram of the MoO3 crystal (Fig. 2, curve a) shows the (0k0) predominant reflection lines which suggest that this compound is of a layered structure packed in the direction of the b-axis. X-ray diffraction patterns obtained on freshly deposited MoO3 films are shown in Fig. 2 (curves b and c).

14

12

S.lo
c

-EG ~4

+50
1
, ,

A,

10

2b 3b 40 5o 6b zo
Angle 28 (degree)

80

Fig. 2. X-ray diffraction diagrams of MoOa. (a) crystalline orthorhombic MoO3 material, (b) thin film grown at T~= 120 C and (c) thin film grown at T~= 250 C

Lithium intercalation in MoO3


3.5 i i I '

i75

10-9
C
'-F
CM

3.0
GJ

/
10-10

+_~2.5
O

"

oo

E
LJ

\
X8
o/0/

=~ 2o
t_J

.,.,.

...... ,....,

.et(1J

1.5

'

i '.

0.0

0.5 1.0 Composifion in LixMO03

1.5

(3J o U .e./C cO
. . . . . Ir

Fig. 3. Discharge curves of Li/MoO3 cells. The MoO3 active cathode material is (a) crystalline e-MoO3 phase, (b) thin film grown at T~= 120 C and (c) thin film grown at T~= 250 C

10-11
c~

/ - - - - ~

,L b A.

/~

The substrate temperature affects widely the structure of MoO3 films, as reported elsewhere [36]. At Ts = 30 C, MoO3 films are competely amorphous and the polycrystalline phase starts to grow for T~> 100 C. Upon increasing the substrate temperature the crystallite size increases and the characteristic crystalline peaks of MoO3 appear. The X-ray diffraction spectrum of films deposited at T~= 120C exhibits both (0k0) and (Okl) orientations representing an e-# mixed phase film (Fig. 2, curve b). We observe that a film deposited at T~= 250 C is rather a polycrystalline material with an orthorhombic structure with a predominant (0k0) orientation (Fig. 2, curve c). The layered structure has been clearly observed by scanning electron microscopy. Figure 3 shows the discharge voltage curves of Li/ MoO3 cells as a function of the lithium concentration in the intercalated active cathode materials. An initial voltage of 3.2 V was measured for MoOa thin-film cathode cells, which is higher than that recorded on the galvanic cell using a crystalline cathode. The measured potentials are in the range 3.2-2.5 V for thin-film batteries (Fig. 3, curves b and c), whereas the range 2.7-1.5 V is obtained with a crystalline cathode (Fig. 3, curve a). Comparing the discharge curves for crystals and films, we may remark that: (i) the cell voltage decreases steadily with the degree of Li insertion in the host materials and reaches a value of 2.45 V for films deposited at T~= 120 C and 2.80 V for films deposited at T~= 250 C (at x = 1.0), (ii) the discharge occurs steadily in the compositional range 0 _ x < 1.5, and (iii) 1.5 electrons are transferred into the host material. The electrochemical process is a classical intercalation mechanism for the lithium insertion [37]. The simultaneous injection of ions and electrons into MoO3 forms a lithium bronze according to the reaction 1.5 Li + + 1.5 e- +MOO3 ~=~Li ts+ (MOO3) 1'5-. (3)

10 -12

(,

'

I,

,'

0.0

0.5 1.0 1.5 Composifion in LixMO0 3

Fig. 4. Chemical diffusion coefficientof lithium ions in LixMoO3 cathodes: (a) crystalline e-MoO3 phase, (b) thin film grown at T~= 120C and (c) thin film grown at T~=250C

It is observed in Fig. 3 that the voltage is lower in cells using MoO3 films deposited at T~= 120 C than in cells with films deposited at T~= 250 C. This result can be attributed to the fact that Li diffusion could be limited by the grain boundary effects which affect the discharge curve. A second possibility is the presence of the mixed a-l? phase which may reduce the standard potential. The discharge curve of the cell fabricated by employing MoO3 films deposited at T~= 250 C (Fig. 3, curve c) is quite stable because of the unique layered structure of c~-MoO3 with large grain size. A third explanation may be oxygen-defects in the host structure involving a lower Fermi level in such a film. Figure 4 shows the chemical diffusion coefficient of Li + ions as a function of the degree of intercalation in LixMoO3 materials. In these experiments (1) is valid: (i) if the duration of the preceding discharge is longer than a critical time to = 6 2/4 D, and (ii) for a limited period of relaxation t < to of the electrochemical cell. In the compositional range 0.2_< x_< 1.2 the values of the chemical diffusion coefficient of lithium varies from 9 x 10-11 t o 1 x 10 9 c m 2 s - 1 in the MoOa crystal. The compositional dependence of D is rather a quadratic function which is due to the nature of the empty sites in the host structure, and the variations of the chemical diffusivity of the intercalated alkali ions can be modeled. In the model of Basu and Worrell [38], the chemical diffusivity is related to the composition D = fix 2 ( 1 - x ) + x ( 1 - x 2 ) , (4)

The reduction process, accompanied by colouration of MoOa, involves the injection of lithium cation and electron into the host according to (3).

where fi is an interaction parameter related to the repulsive interaction energy between alkali ions. According to

176

C. Julien et al.

(4), D shows a maximum at the half-filling site number. Although the uncertainties in the chemical diffusion data shown in Fig. 4 prelude an adequate test of (4), the maximum in the Li~MoO3 data suggests an interaction energy lower than 0.2 eV which is consistent with the value obtained from thermodynamic measurements. The value of the chemical diffusion coefficient in MoO3 thin films is lower than in crystals (Fig. 4, curves b, c). For MoO3 films prepared at Z = 120C a maximum value of 5 x 10 - 1 2 c m 2 s - 1 is obtained at x = 0 . 8 (Fig. 4, curve b). D, which is a function of the composition x, varies according to a quadratic law D = ~x ( 2 - x), (5)

where ~ = 2 x 10 -12 cm 2 s -1. The chemical diffusion coefficient behaviour in MoO3 thin film deposited at 250 C differs from the previous one. We observe that D has an almost constant value of 1.5 x 10 -la cm 2 s -* in the range 0.1 < x < 1.2 and increases at higher concentration of intercalant ions. We tentatively attribute this complex behaviour to the polycrystalline state of MoO3 films grown at high substrate temperature for which the enhancement factor is high. In this case, the intercalation process is partly controlled by the number of ion occupancies in the host lattice of the crystallite in the film. Figure 5 shows the thermodynamic factor, W (in a logarithmic scale), as a function of the degree of intercalation in Li~MoOs materials. In the compositional range 0 . 2 < x < 1.4, W varies from 1.8 to 30 in MoO3 crystals (Fig. 5, curve a). Considering that MoO3 is a layered host for the intercalant, the model of ion-ion interaction can be applied. Armand [39] has proposed such a model to describe the variation of the chemical potential in intercalation compounds. The thermodynamic factor is related to the interaction factor, g,
as

Experimental data are well fitt using (6) which provides an interaction factor 9 = 7.5. This value is of the same order as in TiS2 [40]. The thermodynamic factor of MoO3 thin films varies from 50 to 800 in the compositional range 0 . 2 < x < 1.5. We observe a quasi-linear variation of the composition dependence of W which may be associated with the oxygen-defects in the host lattice. It is a fact that numerous of such defects exist in the film structure, even when the crystallinity is improved by different conditions of preparation. Here, we remark that W is two orders of magnitude higher than the thermodynamic factor in the MoO3 crystal. Using (6) a fit of experimental data is obtained with a high value of the interaction factor ( g = 400). Considering that an MoO3 film is an oxygendeficient material, the model of charge transport in internal defect-materials can be applied [41, 42]. Defects are Li interstitials, Li*, and conduction electrons, e', for example. It has been shown [41] that, in a solid solution where no internal defect reactions occur, the thermodynamic factor is related to the defect concentration (if dilute defects exist) as
W = ( i 0 L i I -~- (i~)e~ 1 ,

(7)

where ~b~-~ * = CLi*/eL~+ and ~bg I = co,/cu+. Equation (7) provides a linear variation of the thermodynamic factor with the degree of intercalation, as shown in Fig. 5 (curve b). The large increase of W may be also associated with the decrease of the electronic mobility in Li~ MoO3 film. For the case in which the chemical diffusion is predominantly determined by ionic species, (2) yields, with the substitution of Do by the conductivity a~ and the thermodynamic factor, for the partial conductivity [43]

a~ to = (qc*D/W) ( F / R T),

(8)

w=

[(1-x)-l+gx].

(6)
10-3
i I l l O n i l i i t

1000 500

l l U l u L n l l ~

m/

"-7-

/
/

u 10-4
/

--,.a
/

lO0
U

/' ,0
/
z

50
u cZP

8 /
/

10-5

'

~z o
E
r-

10
5
/ /

,.2o 10-7

./
/
I I I J

/o

0.0

0.5 1.0 1.5 [.omposifion in LixMoO 3

10-8F , ,, ~
0.0

r--~-~ , , '-

0.5 1.0 1.5 [omposifion in LixNO09

Fig. 5. T h e r m o d y n a m i c factor o f LixMoO3 cathodes: (a) crystalline e - M o O 3 p h a s e a n d (b) thin film g r o w n at T~= 250 C. Dashed lines represent the fit

Fig. 6. Partial ionic conductivityof LixMoO3 cathodes: (a) crystalline e-MoO3 phase and (b) thin film grown at T~= 250 C

Lithium intercalation in MoO3 where to is the transference number te = o-e/(ae + ai). By combining the chemical diffusion coefficient and the thermodynamic factor values, the lithium ion conductivities ai in LixMoO3 materials are shown in Fig. 6. Discharging the cell results in the formation of a mixed conductor; the compound is an electronic and ionic conductor. If the sample is predominantly an electronic conductor, i.e. ao >>ai, the ionic conductivity can be deduced from (8). The conductivity is found to increase with lithium concentration in LixMoO3 crystal to reach about 1.5 x 10-~S c m - I at x = 0.6 (Fig. 6, curve a). One can remark that the ionic transport parameters in LixMoO3 crystal are comparable with those of Lix TiS2 [38]. Figure 6 (curve b) displays the lithium ionic conductivity of MoO3 thin-film cathode. The ionic conductivity is estimated with an average value of 9 x 10 -9 g cm -1 at x = 0.8. We remark that the Li + ionic conductivity is much smaller than for MoO3 crystals. This behaviour may be attributed to the disordered structure of the film. In such a material the conduction paths are not defined as well as they are in the crystalline material because in thin film: (i) the van der Waals plane is not quasi-infinite, (ii) the potential barriers are more important, and (iii) the Li + ions can be trapped by the structural defects. Electrochemical properties of Li~MoO3 materials can be illustrated by an electronic band scheme, as shown in Fig. 7. In the physicists description, the Li/MoO3 cell voltage can be viewed as the difference of the Fermi energies between anode and cathode. In MoO3 there are five O(p~) and three Mo(t2g ) orbitals which interact to form rc and re* bands [44] which form the valence and conduction bands, respectively. The antibonding ~* states hold the extra electrons supplied by the inserted lithium. The narrowing of the conduction band is expected [44] to lead to an increase in the effective electron mass which will affect the position of the Fermi energy in Li~MoO3 phases. In Li-intercalated MoO3 crystals the Fermi energy lies either with the donor level near the bottom of the conduction band or in the conduction band itself (Fig. 7a), whereas in MoO3 thin films the density-of-states might show a band in the middle of the gap and the Fermi level lies in the band. c~-MoO3 is a

177 semiconductor with a large band gap of 3.1 eV. Thus, the highest voltage displayed in thin-film cells can be attributed by the difference between Fermi levels (about 1 eV) in the crystal and thin-film.

3 Conclusion In this work, we have studied the lithium intercalation process in MoO3 host material for two forms of it, the crystalline orthorhombic c~-MoOa and the polyerystalline MoO3 films. The kinetically accessible discharge range amounts to 0_< x_< 1.5 for the Li insertion in MoO3 active cathode materials. In MoO3 crystalline cathodes the chemical diffusion coefficient is 1 x 10 .9 cm 2 s -1 at x = 0 . 7 and the thermodynamic factor reaches 30 at x = 1.4. Values of the chemical diffusion coefficient and thermodynamic factor are similar to those of the Li/TiS2 system. In MoO3 films, the chemical diffusion coefficient ranges from 10 - i t to 10-12cm2s-1 and the thermodynamic factor is two orders of magnitude higher than in MoO3 crystals. This values clearly depict the weaker ionic transport in the disordered phases. However, the cell voltage remains high in films because donating electrons are trapped in the localized band located in the band gap of disordered semiconductors. MoO3 films are interesting as potential electroactive materials for solid-state lithium microbatteries.

References 1. M.S. Whittingham: Science 192, 1126 (1976) 2. D.A. Winn, J.M. Shemilt, B.C.H. Steele: Mater. Res. Bull. 11, 59 (1976) 3. M.S. Whittingham: Prog. Solid State Chem. 12, 41 (1978) 4. K.M. Abraham: J. Power Sources 7, 1 (1981) 5. C. Julien, I. Samaras: Solid State Ionics 27, 101 (1988) 6. A. LeM6haut6, G. Ouvrard, R. Brec, J. Rouxel: Mater. Res. Bull. 12, 1191 (1976) 7. R. Brec: Solid State Ionics 22, 3 (1986) 8. A. J. Jacobsen, L.E. MacCandlish: Solid State Chim. Fr. 29, 355 (1979) 9. A.J. Jacobsen, M.S. Whittingham, S.M. Rich: J. Electrochem. Soc. 126, 887 (1978) 10. M. Armand, L. Coic, P. Palvadeau, J. Rouxel: J. Power Sources 3, 137 (1978) 11. D.W. Murphy, P.A. Christian, J.F. DiSalvo, J.N. Carides: J. Electrochem. Soc. 126, 497 (1979) 12. M.S. Whittingham, M.B. Dines: J. Electrochem. Soc. 124, 1387 (1977) 13. J.B. Goodenough, A. Manthiram, A.C.W.P. James, P. Strobel: In Solid State Ionics, ed. by G.A. Nazri, R.A. Huggins, D.F. Shriver, Vol. 135 (Mater. Res. Soc., Pittsburgh 1989) p. 391 14. L. Campanella, G. Pistoia: J. Electrochem. Soc. 118, 1905 (1971) 15. F.W. Dampier: J. Electrochem. Soc. 121, 656 (1974) 16. N. Margalit: J. Electrochem. Soc. 121, 1460 (1974) 17. J.O. Besenhard, R. Schollhorn: J. Power Sources 1,267 (1976/ 77) 18. J.O. Besenhard, J. Heydecke, H.P. Fritz: Solid State Ionics 6, 215 (1982) 19. J.O. Besenhard, J. Heydecke, E. Wudy, H.P. Fritz: Solid State Ionics 8, 61 (1983)

(a)

(b)

.....

EF

Densitq of sfales
Fig. 7a, b. Schematic representation of the electronic band structure of (a) ~-MoO3 crystal and (b) MoO3 thin film. The energy of the Fermi levelEv lies with the localizedelectronicband upon Li insertion in the host material

178 20. N. Kumagai, N. Kumagai, K. Tanno: J. Appl. Electrochem. 18, 857 (1988) 2l. G.A. Nazri, C. Julien: Solid State Ionics 53-56, 397 (1992) 22. A.T. Jacobson, S.M. Rich: J. Electrochem. Soc. 127, 779 (1980) 23. K. Nassan, D.W. Murphy: J. Non-Cryst. Solids 44, 297 (1981) 24. M.S. Whittingham: J. Electroanal. Chem. 118, 229 (1981) 25. Y. Takeda, R. Kanno, Y. Tsuji, O. Yamamoto : J. Electrochem. Soc. 131, 2006 (1984) 26. Y. Sakurai, J. Yamaki: J. Electrochem. Soc. 132, 512 (1985) 27. M. Wakihara, T. Uchida, T. Morishita, H. Wakamatsu, M. Tanigushi: J. Power Sources 20, 199 (1987) 28. K. Radhakrishnan, B.V.R. Chowdari: Solid State Ionics 51, 197 (1992) 29. N. Miyata, S. Akiyoshi: J. Appl. Phys. 58, 1651 (1985) 30. C.E. Tracey, D.K. Benson: J. Vac. Sci. Technol. A 4, 2377 (1986) 31. S. Ohfuji: Thin Solid Films 115, 299 (1984) 32. O. Bohnke, G. Robert: Solid State Ionics 6, 115 (1982) 33. D. Belanger, G. Laperriere: Chem. Mater. 2, 484 (1990) 34. N. Anwar, C.A. Hogarth: phys. star. sol. (a) 109, 469 (1988) 35. A. Honders, G.H.J. Broers: Solid State Ionics 15, 173 (1985)

C. Julien et al. 36. C. Julien, L. E1-Farh, M. Balkanski, O.M. Hussain, G.A. Nazri: Appl. Surf. Sci. 65--66, 325 (1993) 37. W.R. McKinnon, R.R. Haering: In Modern Aspects of Electrochemistry, ed. by R. White, J.O'M. Bockris, B.E. Conway, Vol. 15 (Plenum, New York 1983) p. 235 38. S. Basu, W.L. Worrell: In Fast Ion Transport in Solids, ed. by P. Vashishta, J.N. Mundy, G.K. Shenoy (North-Holland, Amsterdam 1979)p. 149 39. M.B. Armand: Mat6riaux d'61ectrodes /~ couple redox interne. Dissertation, Grenoble (1978) 40. A. Honders, J.M. der Kinderen, A.H. van Heeren, J.H.W. de Wit, G.H.J. Broers: Solid State Ionics 15, 265 (1985) 41. J. Maier: In Solid State Ionics II, ed. by G.A. Nazri, R.A. Huggins, D.F. Shriver, M. Balkanski, Vol. 210 (Mater. Res. Soc., Pittsburgh 1991) p. 499 42. C. Julien, L. E1-Farh, M. Balkanski, I. Samaras, S.I. Saikh: Mater. Sci. Eng. B 14, 127 (1992) 43. W. Weppner, R.A. Huggins: J. Electrochem. Soc. 124, 1569 (1977) 44. S. Crouch-Baker, P.G. Dickens: Solid State Ionics 32-33, 219 (1989)

Вам также может понравиться