Вы находитесь на странице: 1из 7

Thermochimica Acta 599 (2015) 3541

Contents lists available at ScienceDirect

Thermochimica Acta
journal homepage: www.elsevier.com/locate/tca

Enthalpy of formation of Li1+xMn2xO4 (0 < x < 0.1) spinel phases


Damian M. Cupid *, Alexandra Reif, Hans J. Seifert
Karlsruhe Institute of Technology, Institute for Applied Materials (IAM-AWP), Hermann-von-Helmholtz-Platz 1, Eggenstein-Leopoldshafen 76344, Germany

A R T I C L E I N F O

A B S T R A C T

Article history:
Received 24 June 2014
Received in revised form 31 October 2014
Accepted 2 November 2014
Available online 4 November 2014

The enthalpies of formation of Li1+xMn2xO4 (0 < x < 0.1) spinel phases were measured using high
temperature oxide solution calorimetry in a sodium molybdate solvent at 701  C. One commercially
available LiMn2O4 powder and two Li1+xMn2xO4 samples synthesized using the solgel method were
used for the measurements. The enthalpies of formation of the spinels from the binary constituent oxides
and from the elements become more exothermic with increasing lithium content and increasing average
oxidation state of the manganese cation in the range 0 < x < 0.1. Additionally, the lithium-rich boundary
of the Li1+xMn2xO4 phase for 0 < x < 0.1 was re-investigated using differential thermal analysis
combined with thermogravimetric analysis.
2014 Elsevier B.V. All rights reserved.

Keywords:
Li1+xMn2xO4
Spinel
Oxide solution calorimetry
Enthalpy of formation
Lithium-ion battery

1. Introduction
The lithiummanganeseoxygen system is a key materials
system for the development of advanced cathodes for lithium-ion
batteries. In particular, the Li1+xMn2xO4 cubic spinels (Fd3m
spacegroup) are attractive because they are considered environmentally friendly, are stable against Li extraction, have a high
electrode potential of approximately 4 V versus lithium, and
because the cost of Mn is lower than that of Co which is required for
LiCoO2 production.
The Li1+xMn2xO4 cubic spinels contain a three-dimensional
network of occupied tetrahedral (8a) and vacant interstitial
octahedral (16c) sites. During charge and discharge, the lithium
ions diffuse from the 8a sites to adjacent 16c sites and from these
16c sites to neighboring 8a sites. The reversible capacity of the
cubic spinels originates from the Mn3+/Mn4+ redox pair, in which
the oxidation state of the transition metal cation oscillates
between +3 and +4 during lithium insertion and extraction.
The enthalpies of formation of Li1+xMn2xO4d spinels have
been measured by Wang and Navrotsky [1] using high temperature
oxide solution calorimetry and by Idemoto et al. [2,3] using acid
solution calorimetry. Both Wang and Navrotsky [1] and Idemoto
et al. [2,3] report that the enthalpies of formation of the
Li1+xMn2xO4d spinels become more exothermic with increasing
lithium content (increasing x in Li1+xMn2xO4d up to x = 1/4).
According to Wang and Navrotsky [1], this trend results from the
oxidation of Mn3+ to Mn4+ (an exothermic process) which occurs

* Corresponding author. Tel.: +49 7216 822258; fax: +49 7216 824567.
E-mail address: damian.cupid@kit.edu (D.M. Cupid).
http://dx.doi.org/10.1016/j.tca.2014.11.003
0040-6031/ 2014 Elsevier B.V. All rights reserved.

during Li substitution of Mn on the Mn (16d) sites in the absence of


oxygen vacancy production. Interestingly, Idemoto et al. [2]
showed that an increase in thermodynamic stability of the
Li1+xMn2xO4d cubic spinels, as indicated by a more exothermic
enthalpy of formation, leads to better cycling stabilities and
decreased capacity losses. Generally, however, there is a discrepancy between the enthalpy of formation data of Wang and
Navrotsky [1] and Idemoto et al. [2,3]. For example, Idemoto et al.
[3] show enthalpies of formation of LiMn2O4 from the constituent
oxides of 62.1 1.2, 65.9  1.0, and 70.5  1.3 kJ/mol for
their samples which were annealed at 500  C for 48 h in argon
gas, at 700  C for 24 h in 1 atm O2, and at 500  C for 48 h in
20 atm O2 respectively. Wang and Navrotsky report a value of
82.47  1.79 kJ/mol [1]. In that work, Li2CO3 and MnO2 were
mixed, pelletized, calcined at 500  C in air for 12 h and
subsequently annealed at 750  C for 24 h in air to produce
LiMn2O4. Both Wang and Navrotsky [1] and another work of
Idemoto et al. [2] measured the enthalpy of formation of
Li4/3Mn5/3O4 from the constituent oxides. The value of Idemoto
et al. [2] is 101.7  1.5 kJ/mol whereas that of Wang and
Navrotsky is 110.26  2.60 kJ/mol. It must be mentioned that
Wang and Navrotsky [1] measured the oxygen content in nominal
Li4/3Mn5/3O4 (their reported composition was Li4/3Mn5/3O3.849) and
took the deviation from stoichiometric oxygen contents into
account in the thermodynamic cycle.
The aim of this work is to clarify the discrepancies in the
published enthalpies of formation data from Wang and Navrotsky
[1] and Idemoto et al. [2,3]. We therefore independently measured
the enthalpies of formation of the Li1+xMn2xO4 spinels using high
temperature oxide melt drop solution calorimetry in a newly
installed TianCalvet high temperature calorimeter. Additionally,

36

D.M. Cupid et al. / Thermochimica Acta 599 (2015) 3541

the Li-rich phase boundary of the Li1+xMn2xO4 spinel phase at


ambient oxygen partial pressures was re-investigated using
differential thermal analysis combined with thermogravimetric
analysis. These results further resolve contradictions between
Paulsen and Dahn [4] and Luo and Martin [5] in the location of the
Li-rich phase boundary of the Li1+xMn2xO4 phase in air. These
contradictions are already detailed in Cupid et al. [6], in which the
Li-rich phase boundary proposed by Luo and Martin [5] was
conrmed. The thermochemical and phase diagram data presented
in this work are valuable input data for the development of
CALPHAD-based thermodynamic descriptions of the LiMnO
system, which can eventually be used to calculate open
circuit voltages and coulometric titration curves based on the
Gibbs free energy descriptions of the assessed phases in the
multi-component systems.
2. Experimental
The Li1+xMn2xO4 samples were prepared by the solgel
method using Li-acetate, Mn-acetate and adipic acid. The origin
and purities of the starting chemicals are listed in Table 1. The
required amounts of Li-acetate, Mn-acetate and adipic acid were
dissolved separately in distilled water. The amount of adipic acid
used was chosen to yield a 1:1 ratio of total metal ions to adipic
acid in the nal solution. The solutions were mixed at room
temperature and ammonium hydroxide was added to attain a pH
value of 6.5. The resulting solution was heated at 80  C and stirred
continuously until a colorless transparent gel was formed. The gel
was subsequently dried for 48 h in a drying oven at 120  C. The
dried gel was ground into a ne powder, placed in an alumina
crucible and pre-calcinated at 350  C for 5 h in air. A 3  C/min
heating rate to 350  C was used, which ensures that a sudden
evolution of CO, CO2, and H2O gases does not take place during the
highly exothermic combustion of the organic components. The
precursor powders were furnace cooled after the rst heat
treatment, ground into a ne powder and heat treated a second
time for 15 h in air at the required temperature. At the end of the
second heat treatment, the samples were taken out of the oven
and cooled in air. XRD was performed on the samples to conrm
the presence of Li1+xMn2xO4. In addition to the self-prepared
Li1+xMn2xO4 samples, a commercially available LiMn2O4 sample
was also investigated in this work. The origin and purity of the
commercial LiMn2O4 sample are given in Table 1.
Differential thermal analysis was performed using a Setaram
TAG high performance symmetrical system (Setaram Instrumentation, Caluire, France). In all measurements, approximately 20 mg
samples which were pressed using a 3 mm diameter die were

investigated. In the rst set of experiments, samples which were


subjected to the rst heat treatment at 350  C were cycled three
times successively between 150 and 1000  C with a heating rate of
10  C/min in an Ar/20 volume% O2 gas mixture (origin and purity
in Table 1) to determine the sample phase transformation
temperatures. The temperature for the second heat treatment
was determined based on the thermal analysis results to ensure
the presence of only the Li1+xMn2xO4 phase after the heat
treatment. In the second set of thermal analysis experiments, the
heat-treated samples were cycled between 150 and 1200  C in the
Ar/20 volume% O2 gas mixture with a heating rate of 10  C/min. All
measurements were performed in platinum crucibles.
The compositions of the samples after the second heat
treatment and that of the commercial sample were measured
using inductively coupled plasma with optical emission spectroscopy (ICP-OES). 5 mg samples were used (weighing accuracy
 0.002 mg), which were dissolved into 3 mL hydrochloric acid at
80  C. A constant amount of an yttrium and sodium solution was
added to serve as an internal standard for the manganese and
lithium measurements, respectively. The rest solution was made
up with de-ionized water up to a volume of 500 ml. An OPTIMA
4300 DV from PerkinElmer with an echelle spectrometer equipped
with segmented-array charge-coupled detector, which detects the
emission lines for the analytic solution and the internal standard
simultaneously, was used for the ICP-OES measurements.
High temperature oxide melt drop solution calorimetry was
performed using a Setaram AlexSys-1000 TianCalvet calorimeter
(Setaram Instrumentation, Caluire, France). The AlexSys-1000 is a
commercial TianCalvet calorimeter constructed according to the
specications of the custom-built calorimeters as detailed in
Navrotsky [7,8]. This calorimeter is suitable for measuring
enthalpies of drop solution in oxide melts. In the course of this
work, the AlexSys-1000 was installed in a dedicated room kept at
constant temperature in which there are no rapid air ows. These
conditions are required to ensure stability of the instrument heat
ow signal (baseline) during the course of the measurements. The
set-point temperatures of the three-zone furnace were adjusted to
maintain a at temperature prole with an average temperature of
701  C in the platinum crucibles of 10 cm length inside the
calorimeter. Both sides of the twin calorimeter were calibrated
independently using 1.5 mm diameter sapphire spheres of
approximately 7 mg mass (origin and purity in Table 1), which
were dropped from room temperature into the empty platinum
crucibles in the calorimeter at 701  C.
Sodium molybdate (3Na2O4MoO3) was used as the solvent for
the high temperature drop solution calorimetry. The sodium
molybdate solution was prepared by thoroughly mixing

Table 1
Sample table.
Chemical name

Source

Initial mass fraction purity

Purication method

Lithium acetate di-hydratea


Manganese(II) acetate tetrahydrateb
Adipic acidc
Sodium molybdate dihydrated
Moylbdenum(VI) oxidee
LiMn2O4
Manganese(IV) oxidef
Sapphire spheres
Argon
Ar/20 volume% O2

Alfa Aesar
AppliChem
Merck Schuchardt OHG
Merck KGaA
Merck KGaA
MTI
Alfa Aesar
Alfa Aesar
Air Liquide
basi Schberl GmbH & Co

0.99
0.99
0.99
0.995
0.99
0.995
0.99997
0.9999
0.9999g
0.1985O2h

none
none
none
none
none
none
none
none
none
none

a
b
c
d
e
g
h

LiOOCCH32H2O.
(CH3COO)2Mn4H2O.
C6H10O4.
Na2MoO42H2O.
MoO3.fMoO2.
Purity expressed in mole fraction.
The concentration of O2 in Ar/20 volume% O2 is given in volume fraction. The standard uncertainty is u = 0.0040.

D.M. Cupid et al. / Thermochimica Acta 599 (2015) 3541

37

Table 2
Expected and measured compositionsa of the Li1+xMn2xO4 spinel samples. The lithium and manganese compositions were measured using ICP-OES. The Li1+xMn2xO4
(x = 0.1 and 0.2) samples were prepared by the solgel method using the starting compounds Li-acetate, Mn-acetate, and adipic acid. The origin and purities of the starting
chemicals are given in Table 1.
Nominal stoichiometry

LiMn2O4
Li1.1Mn1.9O4
Li1.2Mn1.8O4

Measured data

Calculated stoichiometry

Li mass%

Mn mass%

Li

Mn

Li/Mn

4.00  0.01
4.34  0.07
4.81  0.01

60.38  0.07
60.39  0.66
58.10  0.07

1.032  0.002
1.088  0.019
1.188  0.003

1.968  0.002
1.912  0.019
1.812  0.003

0.524  0.002
0.569  0.013
0.656  0.002

The number following the  symbol is the expanded uncertainty U, calculated as U = ku for the measured data and U = kuc for the calculated stoichiometries, where u and uc
are the standard and combined uncertainties respectively. A coverage factor k = 2 was used, corresponding to a condence interval of 95 percent.

appropriate amounts of sodium molybdate dihydrate


(Na2MoO42H2O) and molybdenum (IV) oxide so that the nal
molar ratio between Na2O and MoO3 is 3:4. The origin and purities
of sodium molybdate dihydrate and molybdenum(IV) oxide are
given in Table 1. The mixture was heated at 700  C in air for 5 h,
furnace cooled, and then ground into a powder. Approximately
28 g of the sodium molybdate powder was weighed into each
platinum crucible and used for the solution calorimetry.
The enthalpy of drop solution of Mn2O3 in sodium molybdate
was measured to validate our calibration factors and to check our
results against data from Wang and Navrotsky [1]. The Mn2O3
samples were produced by annealing high purity MnO2 (origin and
purity in Table 1) at 700  C for 24 h in air. After the heat treatment,
the sample was taken out of the oven and cooled in air. XRD
performed on the sample conrmed the presence of Mn2O3
after the heat treatment. After the enthalpy of drop solution of
Mn2O3 was determined, the enthalpies of drop solution of the
Li1+xMnxO4 compounds were then measured.
For the high temperature oxide melt drop solution measurements, all samples (Mn2O3 and Li1+xMn2xO4) were pressed into
pellets with a diameter of 3 mm and dropped from room
temperature into the sodium molybdate solution at 701  C. The
masses of the pellets were between 5 and 6 mg (weighing
accuracy  0.005 mg). Ar gas with a ow rate of 40 ml/min was
used as a ushing gas to maintain a constant environment
above the sodium molybdate solution and Ar gas with a ow rate of
5 ml/min was bubbled continuously into the solution to stir the
solution and prevent local solvent saturation. The origin and purity
of the Ar gas used is given in Table 1. The time for one drop was

Fig. 1. Differential thermal analysis heat ow signal of Li1.2Mn1.8O4 measured using


a heating rate of 10  C/min in an Ar/20 volume% O2 gas mixture. The rst heating
cycle is shown as the dashed line whereas the subsequent three heating and cooling
cycles are shown as full lines.

approximately 45 min and drops were performed every 2 h once


the baseline was observed to be stable between drops.
3. Results
3.1. Stoichiometry of Li1+xMn2xO4
The lithium and manganese contents of the samples were
determined by ICP-OES. The nominal and measured compositions
are given in Table 2. The numerical values of x in Li1+xMn2xO4
were calculated from the measured lithium n(Li) and manganese n
(Mn) ion concentrations and by assuming the ratio n(Li)/n
(Mn) = (1 + x)/(2  x). The nominal stoichiometries and those
calculated from the results of the ICP-OES analyses are in good
agreement with each other.
3.2. Lithium-rich phase boundary of Li1+xMn2xO4
The lithium-rich boundary of the Li1+xMn2xO4 phase in Ar/20
volume% O2 gas was conrmed using differential thermal analysis
coupled with thermogravimetric analysis. The heat ow signal for
Li1.2Mn1.8O4 and the commercial LiMn2O4 sample are given in
Figs. 1 and 2, respectively. In both samples, a phase transformation
can be detected on rst heating which does not appear on cooling
and on subsequent heating and cooling cycles. This phase
transformation is associated with the reaction:
y
Li1x Mn2x O4d $zLi2 MnO3 Li1x2z Mn2xz O4d3zy O2
2
during which small amounts of Li2MnO3 are formed accompanied
with a shift in the composition of the cubic spinel phase and
simultaneous evolution of oxygen gas. This reaction has been

Fig. 2. Differential thermal analysis heat ow signal of commercially available


LiMn2O4 measured using a heating rate of 10  C/min in an Ar/20 volume% O2 gas
mixture. The rst heating cycle is shown as the dashed line whereas the subsequent
two heating and cooling cycles are shown as full lines.

38

D.M. Cupid et al. / Thermochimica Acta 599 (2015) 3541

Fig. 3. Thermogravimetric analysis results of Li1.2Mn1.8O4 measured using a heating


rate of 10  C/min in an Ar/20 volume% O2 gas mixture. The rst heating cycle is
shown as the dashed line whereas the subsequent two heating and cooling cycles
are shown as full lines.

discussed extensively in the literature by Thackeray et al. [9], Tsuji


et al. [10], and Luo and Martin [5].
In this work, the mass loss associated with oxygen gas evolution
was measured using simultaneous thermogravimetric analysis. The
results for self- prepared Li1.2Mn1.8O4 and commercial LiMn2O4 are
given in Figs. 3 and 4, respectively. In both samples, an irreversible
mass loss on rst heating is indicated, followed by reversible and
near-reversible mass losses on subsequent heating and cooling
cycles for Li1.2Mn1.8O4 and commercial LiMn2O4 respectively. The
rst derivatives of the mass changes with respect to time during the
rst heating cycle are calculated and shown against sample
temperature in Figs. 5 and 6 for Li1.2Mn1.8O4 and commercial
LiMn2O4, respectively. The rst deviation of the rst derivatives from
the steady state baseline can be used to assign the temperature at
which evolution of O2 gas and formation of Li2MnO3 occurs. The
results of these investigations are given in Table 3. This temperature
indicates the location of the lithium-rich boundary of the
Li1+xMn2xO4 cubic spinel phase at the respective sample

Fig. 4. Thermogravimetric analysis of commercially available LiMn2O4 measured


using a heating rate of 10  C/min in an Ar/20 volume% O2 gas mixture. The rst
heating cycle is shown as the dashed line whereas the subsequent three heating and
cooling cycles are shown as full lines.

Fig. 5. Determination of the lithium-rich boundary of the Li1+xMn2xO4 phase


using the rst derivative of the mass change signal with time for Li1.2Mn1.8O4.

composition. The temperatures measured in this work are


compared to other literature data from Gao and Dahn [11], Luo
and Martin [5], and Cupid et al. [6] in Fig. 7. A very good agreement is
shown with the data of Luo and Martin [5] and Cupid et al. [6]. Based
on these results, nal heat treatment temperatures were 600 and
650  C for the Li1.2Mn1.8O4 and Li1.1Mn1.9O4 samples respectively.
Other phase transformations which were detected at higher
temperatures are also summarized in Table 3. Although the exact
nature of the phase transformations were not investigated in this
work, they are well known and have already been characterized by
Thackeray et al. [9] and Tsuji et al. [10]. According to the LiMnO
phase diagram in air from Paulsen and Dahn [4], the reaction at T2
is associated with the formation of LiMnO2 from tetragonal spinel
Li1xMn2+xO4 and Li2MnO3 whereas the reaction at T3 is
associated with the decomposition of tetragonal Li1xMn2+xO4
to form LixMn3xO4 and additional LiMnO2. Since the T2 reaction
temperature is, within experimental error, independent of the
starting composition of the Li1+xMn2xO4 phase, the reaction at T2
can be considered an invariant four phase reaction in the LiMnO
system between LiMnO2, Li2MnO3, tetragonal spinel Li1xMn2+xO4
and oxygen at a partial pressure of p(O2) = 0.2 atm. In fact,
Thackeray et al. [9] and Tsuji et al. [10] both reported T2
temperatures of 930 and 937  C respectively, which are similar

Fig. 6. Determination of the lithium-rich boundary of the Li1+xMn2xO4 phase


using the rst derivative of the mass change signal with time for commercially
available LiMn2O4

D.M. Cupid et al. / Thermochimica Acta 599 (2015) 3541


Table 3
Phase transformation temperaturesa of the one commercial (LiMn2O4) and two selfprepared Li1+xMn2xO4 samples in an Ar/20 volume% O2 gas mixture at p = 0.1 MPab .
Nominal stoichiometry

LiMn2O4
Li1.1Mn1.9O4
Li1.2Mn1.8O4

Transformation temperatures ( C)
T1

T2

T3

757  3
691  3
612  3

934  1
932  1
932  1

1047  4
1062  2
1055  1

The standard uncertainties for T1 are u(T1) = 3  C and are assigned based on the
uncertainty in selecting the temperature at which the rst derivative of the mass
change with time deviates from the steady-state baseline as shown in Figs. 5 and 6.
The number following the  symbol for the T2 and T3 temperatures is the expanded
uncertainty U, calculated as U = ku where u is the standard uncertainty and a
coverage factor k = 2 was used, corresponding to a condence interval of 95 percent.
b
The standard uncertainty in p is u(p) = 5 kPa.
a

to our results. The T3 temperature reported in this work depends


on the composition of the sample investigated. Thackeray et al. [9]
reported a temperature of 1060  C whereas Tsuji et al. [10] reported
a temperature of 1077  C, both for stoichiometric LiMn2O4. The
large scatter in the data may indicate that the kinetics of
decomposition of tetragonal Li1xMn2+xO4 with formation of
LixMn3xO4 and LiMnO2 may play a signicant role in the course
of the reaction.

39

Table 4
Enthalpies of drop solutiona at p = 0.1 MPab of the samples measured in this work.
Approximately 28 g of sodium molybdate (3Na2O4MoO3) solvent was used for the
high temperature oxide melt drop solution calorimetry. The solvent was maintained
at a temperature of 701  C inside the platinum crucibles in the calorimeter. Each
sample weighed between 5 and 6 mg and the recorded heat effects for each sample
were between 5 and 10 J. The enthalpy of drop solution is the sum of the enthalpy
increment from 298 to 974 K and the enthalpy of dissolution of the samples in
sodium molybdate at 974 K.
Nominal stoichiometry

Enthalpy of drop solution [kJ/mol] Number of drops

Mn2O3
LiMn2O4
Li1.1Mn1.9O4
Li1.2Mn1.8O4

165.32  1.61
249.25  2.08
253.71  2.53
261.28  1.78

8
14
11
12

a
The number following the  symbol is the expanded uncertainty U, calculated
as U = ku where u is the standard uncertainty and a coverage factor k = 2 was used,
corresponding to a condence interval of 95 percent.
b
The standard uncertainty in p is u(p) = 5 kPa

Li1+xMn2xO4 compounds measured in this work determined by a


least squares regression method.
The reaction equations and thermodynamic cycles used to
calculate the enthalpy of formation of Li1+xMn2xO4 from the
constituent oxides are as follows:
Li2O(s,

298 K) ! Li2O(sol, 298 K)

DHdsLi2O

3.3. High temperature oxide solution calorimetry


Mn2O3(s,
The enthalpy of drop solution of Mn2O3 and the Li1xMn2+xO4
compounds are given in Table 4 in kJ/mol. These results are based
on the heat effects which were measured in the range from 5 to 10 J
for each drop. The enthalpy of drop solution for the commercial
and self-prepared Li1+xMn2xO4 compounds are compared to the
work of Wang and Navrotsky [1] in Fig. 8. The solid line in Fig. 8 is
the line of best t for the enthalpies of drop solution of the three

MnO2(s,

298 K) ! Mn2O3(sol, 298 K)

298 K) ! MnO2(sol, 298 K)

Li1+xMn2xO4(s,

DHdsMn2O3

DHdsMnO2

298 K) ! Li1+xMn2xO4(sol, 298 K)

DHdsLi1+xMn2xO4

1 x
1  3x
 Li2 O
 Mn2 O3
2
2
1 2x  MnO2 ! Li1x Mn2x O4

DHfox Li1x Mn2x O4

1 x
1  3x
 DHds Li2 O
 DHds Mn2 O3
2
2
1 2x  DHds MnO2  DHds Li1x Mn2x O4

DHfox Li1x Mn2x O4

Fig. 7. Thermogravimetric analysis results for the commercial and two


self-prepared Li1+xMn2xO4 samples measured using a heating rate of 10  C/min
in an Ar/20 volume% O2 gas mixture with the lithium-rich phase boundary of the
Li1+xMn2xO4 phase from previous works.

Fig. 8. Enthalpy of drop solution for one commercial (lled triangle) and two
self-prepared (lled diamonds) Li1+xMn2xO4 spinels compared to the data of
Wang and Navrotsky (open squares) [1].

40

D.M. Cupid et al. / Thermochimica Acta 599 (2015) 3541

Table 5
Enthalpy of formationa of LiMnO compounds measured using high temperature oxide melt drop solution calorimetry at p = 0.1 MPab.
Nominal
stoichiometry

Enthalpy of formation from the


oxides DHfox (kJ/mol)

Enthalpy of formation from the oxides

DHfox (kJ/mol of atoms)

Enthalpy of formation from the


elements DHfele (kJ/mol)

Enthalpy of formation from the


elements DHfele (kJ/mol of atoms)

LiMn2O4
Li1.1Mn1.9O4
Li1.2Mn1.8O4

85.36  2.67
91.86  3.06
103.09  2.57

12.19  0.38
13.12  0.44
14.73  0.37

1381.65  2.85
1382.88  3.24
1384.64  2.83

197.38  0.41
197.55  0.46
197.81  0.40

a
The number following the  symbol is the expanded uncertainty U, calculated as U = kuc where uc is the standard combined uncertainty and a coverage factor k = 2 was
used, corresponding to a condence interval of 95 percent.
b
The standard uncertainty in p is u(p) = 5 kPa.

The enthalpies of formation of the Li1+xMn2xO4 spinel


phases from the binary oxides are given in Table 5 and plotted
in Fig. 9. Data for the enthalpies of drop solution of
Li2O (DHdsLi2O = 93.02  2.24 kJ/mol) and MnO2 (DHdsMnO2 =
128.92  0.91 kJ/mol) used to calculate the enthalpies of
formation of the ternary compounds from the binary oxides
were taken from [12] and [1] respectively, whereas the enthalpy
of drop solution of Mn2O3 (DHdsMn2O3 = 128.92  0.91 kJ/mol)
was taken from this work. The solid line shown in Fig. 9 is the
line of best t determined using a least squares regression method
for the enthalpies of formation of the three Li1+xMn2xO4
compounds from the binary constituent oxides which were
measured in this work. Our results show a very good agreement
with the data of Wang and Navrotsky [1], even for the commercial
sample.
The enthalpies of formation of the Li1+xMn2xO4 compounds
from the elements are given in Table 5 and are compared to the
data of Wang and Navrotsky [1] in Fig. 10. The enthalpies of
formation of the compounds from the elements were calculated
using the standard enthalpies of formation of the binary
constituent oxides from Glushko et al. [13] (597.935 
0.334 kJ/mol for Li2O, 521.493  0.836 kJ/mol for MnO2 and
957  0.836 kJ/mol for Mn2O3). These values are only
slightly different to those given in Robie and Hemingway
(597.9  2.1 kJ/mol for Li2O, 520  0.7 kJ/mol for MnO2 and
959  1 kJ/mol for Mn2O3) [14], which were used by Wang and
Navrotsky [1]. The solid line in Fig. 10 is the line of best t using a
least squares regression method for the enthalpies of formation of
the three Li1+xMn2xO4 compounds from the elements which
were measured in this work.

Fig. 9. Enthalpy of formation from the oxides for one commercial (lled triangle)
and two self-prepared (lled diamonds) Li1+xMn2xO4 spinels compared to the data
of Wang and Navrotsky (open squares) [1].

4. Discussion
The enthalpies of formation of the Li1+xMn2xO4 spinels
measured in this work are in better agreement with the work of
Wang and Navrotsky [1] than with that of Idemoto et al. [3]. For
example, our value for the enthalpy of formation of Li1.1Mn1.9O4
from the binary oxides is 91.86  3.06 kJ/mol. Wang and
Navrotsky [1] report a value of 91.83  2.55 kJ/mol whereas
Idemoto et al. [3] show values of 67.5  1.2, 82.6  1.4, and
93.6  1.0 kJ/mol for samples which were annealed in Ar, 1 atm
O2, and 20 atm O2, respectively. Additionally, we report a value
for the enthalpy of formation from the constituent binary
oxides for Li1.03Mn1.97O4 (the commercial sample with the
nominal composition LiMn2O4) of 85.36  2.67 kJ/mol, whereas
Idemoto et al. [3] report the values 64.1 1.9, 69.1 1.2, and
77.3  1.3 kJ/mol for the sample with composition Li1.03Mn1.97O4
annealed in Ar, 1 atm O2, and 20 atm O2 respectively. Although
Wang and Navrotsky [1] did not measure a sample of this
composition, their values for the LiMn2O4 and Li1.05Mn1.95O4
samples indicate that the enthalpy of formation for Li1.03Mn1.97O4
from the binary oxides should be between 82.47  1.79 (for
LiMn2O4) and 89.92  2.06 kJ/mol (for Li1.05Mn1.95O4). However,
the large deviation in the measured enthalpies of formation for
samples of the same composition but heat treated at different
conditions as detailed in Idemoto et al. [3] show that the resulting
crystal chemistry inuences the measured enthalpy of formation.
Generally, the enthalpies of formation of the Li1+xMn2xO4
(0 < x < 0.1) spinels from the binary constituent oxides and from
the elements decrease with increasing lithium content as shown
in Figs. 9 and 10, respectively. The spinel compounds are more
stable than their constituent binary oxides and elements, and this

Fig. 10. Enthalpy of formation from the elements for one commercial (lled
triangle) and two self-prepared (lled diamonds) Li1+xMn2xO4 spinels compared
to the data of Wang and Navrotsky (open squares) [1].

D.M. Cupid et al. / Thermochimica Acta 599 (2015) 3541

41

to our work, the values of Wang and Navrotsky [1] are more
reliable. The enthalpies of formation of the Li1+xMn2xO4 spinels
from the binary oxides and from the elements for the composition
range 0 < x < 0.1 become more exothermic with increasing lithium
content. The main factor inuencing the enthalpy of formation of
the spinels from the binary oxides could be the average oxidation
state of the manganese cation. Additionally, the lithium-rich
boundary of the Li1+xMn2xO4 cubic spinel phase was once
again conrmed using differential thermal analysis combined
with thermogravimetric analysis. These phase diagram and
thermochemical data are essential for the advanced thermodynamic modeling of the LiMnO system using the CALPHAD
method. The thermochemical and phase stability data presented in
this work could eventually be combined with electrochemical
testing to make clear statements on the cycling stabilities of
electrode materials.
Acknowledgements
Fig. 11. Enthalpy of formation from the oxides against average manganese
oxidation state of one commercial (lled triangle) and two self-prepared (lled
diamonds) Li1+xMn2xO4 spinels compared to the data of Wang and Navrotsky
(open squares) [1].

relative stability increases with increasing lithium content. Wang


and Navrotsky separated the Li1+xMn2xO4 compounds into the
oxygen stoichiometric (x < 0.2) and oxygen nonstoichiometric
(x  0.25) spinels. The enthalpy of formation of the oxygen
stoichiometric spinels decreases with increasing x due to
reduction of Mn3+ to Mn4+. On the other hand, the endothermic
formation of oxygen vacancies to maintain charge neutrality in
the oxygen non-stoichiometric spinels results in a positive
deviation from this trend. When the enthalpies of formation of
the Li1+xMn2xO4 spinels in this work and from that of Wang and
Navrotsky [1] are plotted against the average oxidation state of
manganese, a linear trend is observed (Fig. 11). The oxygen
non-stoichiometric spinels with average manganese oxidation
states between 3.779 and 3.819 deviate only slightly from this
linear trend. Generally, however, the linear behavior indicates that
the enthalpy of formation of the Li1+xMn2xO4 spinels from the
constituent oxides is primarily determined by the average
oxidation state of the Mn cation. For lithium contents x < 0.2,
the average oxidation state of the Mn cation is controlled solely
by the amount of lithium substitution on Mn sites. At higher
lithium substitutions (x  0.25) the generation of oxygen
vacancies further contributes to the nal average oxidation state
of the Mn cation [1]. Since Idemoto et al. [2] show that the cycling
stability of the spinel electrodes improves with more exothermic
enthalpies of formation, however, one can assume that increasing
the amount of Mn4+ cations in the spinel structure improves
electrode stability.
5. Conclusion
In this work, the enthalpies of formation of Li1+xMn2xO4
spinels were measured using high temperature oxide solution
calorimetry to clarify the discrepancy in the literature between the
data of Wang and Navrotsky [1] and Idemoto et al. [2,3]. According

The authors would like to thank Dr. Thomas Bergfeldt for the
ICP-OES measurements and Dr. Johannes Prll for providing the
commercial powder and the intensive discussions. This work was
funded by the German Research Foundation (DFG) Priority
Program 1473 WeNDeLIB Materials with New Design for
Improved Lithium Ion Batteries and the Helmholtz Association
through the Young Investigator Group VH-NG-1057.
References
[1] M.J. Wang, A. Navrotsky, Thermochemistry of Li1+xMn2xO4 (0  x  1/3)
spinel, J. Solid State Chem. 178 (4) (2005) 11821189.
[2] Y. Idemoto, T. Ozasa, N. Koura, Thermodynamic investigation and cathode
performance of LiMnO spinel system as cathode active material for lithium
secondary battery, J. Ceram. Soc. Jpn. 109 (9) (2001) 771776.
[3] Y. Idemoto, S. Ogawa, Y. Uemura, N. Koura, Thermodynamic stability and
cathode performance of Li1+xMn2xO4 as a cathode active material for lithium
secondary battery, J. Ceram. Soc. Jpn. 108 (9) (2000) 848853.
[4] J.M. Paulsen, J.R. Dahn, Phase diagram of LiMnO spinel in air, Chem. Mater. 11
(11) (1999) 30653079.
[5] C.H. Luo, M. Martin, Stability and defect structure of spinels Li1+xMn2xO4delta: I. In situ investigations on the stability eld of the spinel phase, J. Mater.
Sci. 42 (6) (2007) 19551964.
[6] D.M. Cupid, T. Lehmann, T. Bergfeldt, H. Berndt, H.J. Seifert, Investigation of the
lithium-rich boundary of the Li1+xMn2xO4 cubic spinel phase in air, J. Mater.
Sci. 48 (9) (2013) 33953403.
[7] A. Navrotsky, Progress and new directions in high-temperature calorimetry,
Phys. Chem. Miner. (1977) 89104.
[8] A. Navrotsky, Progress and new directions in high temperature calorimetry
revisited, Phys. Chem. Miner. 24 (3) (1997) 222241.
[9] M.M. Thackeray, M.F. Mansuetto, D.W. Dees, D.R. Vissers, The thermal stability
of lithiummanganeseoxide spinel phases, Mater. Res. Bull. 31 (2) (1996)
133140.
[10] H. Tsuji, Thermodynamic properties of undoped and Fe-doped LiMn2O4 at high
temperature, J. Phys. Chem. Solids (2005) 283287.
[11] Y. Gao, J.R. Dahn, The high temperature phase diagram of Li1+xMn2xO4 and its
implications, J. Electrochem. Soc. 143 (6) (1996) 17831788.
[12] M.J. Wang, A. Navrotsky, Enthalpy of formation of LiNiO2, LiCoO2 and their
solid solutions LiNi1xCOxO2, Solid State Ionics 166 (12) (2004) 167173.
[13] V.P. Glushko, V.A. Medvedev, L.V. Gurvich, Thermal Constants of Substances,
John Wiley, New York, 1999.
[14] R.A. Robie, B.S. Hemingway, Thermodynamic Properties of Minerals and
Related Substances at 298.15 K and 1 bar (105 Pa) Pressure and at
Higher Temperatures, United States Government Printing Ofce, Washington,
1992.

Вам также может понравиться