Вы находитесь на странице: 1из 16

J Mater Sci (2012) 47:20572071

DOI 10.1007/s10853-011-5918-5

REVIEW

Synthesis and applications of molybdenum (IV) oxide


Caleb A. Ellefson Oscar Marin-Flores

Su Ha M. Grant Norton

Received: 11 May 2011 / Accepted: 30 August 2011 / Published online: 16 September 2011
Springer Science+Business Media, LLC 2011

Abstract Molybdenum dioxide (MoO2) is a transition of its weakly bonded layered crystal structure, similar to
metal oxide with unusual metal-like electrical conductivity graphite. High temperature furnace elements are commonly
and high catalytic activity toward reforming hydrocarbons. made of molybdenum disilicide (MoSi2).
This review covers the synthesis techniques used to fabri- Although molybdenum has oxidation states ranging from
cate MoO2 in a variety of morphologies and particle sizes. ?2 to ?6, oxides exist mainly in two forms: molybdenum
Processing from molybdenum ore and reduction from (IV) and molybdenum (VI) oxide. Hexavalent MoO3 is the
MoO3 are also covered, with emphasis on reduction principal product of MoS2 roasting, and is also the main
mechanisms and kinetic considerations. Discussions of molybdenum compound added in steel production. It is also
various solution-based and gas phase synthesis techniques of interest as a semiconductor [1], a field-emitter [2], an
shed light on strategies to achieve various unique mor- electrochomic material [3, 4], photochromic material [5],
phologies, which leads into a brief discussion of nanoscale and a gas sensor [68]. Tetravalent MoO2 exhibits metal-
MoO2. Nanoscale MoO2 is of interest for important tech- like electronic conductivity [9] because of the existence of
nological applications including catalysts for partial oxi- delocalized electrons in its valence band [10]. Recently,
dation of hydrocarbons, solid oxide fuel cell anodes, and MoO2 has garnered much interest for rechargeable lithium
high-capacity reversible lithium ion battery anodes. ion battery (LIB) anodes [e.g., 9]. Both MoO3 and MoO2
have been used extensively as a catalyst in hydrocarbon
reforming processes [e.g., 1113]. In particular, MoO2 has
Introduction been shown to be coke resistant [14] and sulfur tolerant [15]
during the partial oxidation of hydrocarbons. Because of its
Molybdenum may be most well known as an alloying electronic conductivity, MoO2 also has potential as an
element in various steels, such as 316 stainless steel and anode material for fuel flexible solid oxide fuel cells [e.g.,
most high strength steels, where it is used to improve 16, 17]. This review focuses on a comparison of synthesis
corrosion resistance. However, molybdenum compounds methods of MoO2 and an overview of current research
have long been in use for a variety of applications. efforts in specific application areas.
Molybdenum disulfide, MoS2, is a solid lubricant because

Molybdenum (IV) oxide synthesis techniques


C. A. Ellefson  M. G. Norton (&)
School of Mechanical and Materials Engineering, Washington
This section will present a review of MoO2 synthesis
State University, P.O. Box 642920, Pullman, WA 99164-2920,
USA techniques. First, the mineral processing of molybdenum
e-mail: norton@mme.wsu.edu ores is briefly considered. Then, direct hydrogen reduction
of MoO3 will be discussed in detail. We will describe
O. Marin-Flores  S. Ha
hydrothermal and ambient temperature solution based
Voiland School of Chemical Engineering and Bioengineering,
Washington State University, P.O. Box 642710, Pullman, methods, non-solution methods, and because of special
WA 99164-2710, USA interest, research into nanoscale MoO2 will be reviewed.

123
2058 J Mater Sci (2012) 47:20572071

Finally, morphologically unique MoO3, a precursor for 550 C. Much research has been dedicated to the thermo-
MoO2 synthesis, will be highlighted. dynamics and kinetics of this reaction. Hagg and Magneli
performed the first phase studies of molybdenum oxides in
Mineral processing 1944 [19]. Later, Magneli and co workers [20, 21] detailed
the MoO2 crystal structure. Kihlborg [22, 23] identified
Molybdenum is found in the earths crust primarily as the several intermediate oxides with valence states between 4?
mineral molybdenite (MoS2). After crushing, grinding, flo- and 6?. Early research simplified the direct reduction
tation, and leaching of mined ore, molybdenite concentrate reaction to a simple one step full reduction (Eq. 1) with
is roasted in air at 500650 C yielding technical intermediate phases playing an unimportant role, and
molybdenum oxide, an impure product composed primarily appearing under mild conditions (\450 C) with slow
of MoO3 [18]. Many laboratory and industrial applications reaction kinetics [24].
require further purification of oxide products. Common Full reduction:
purification methods performed on an industrial scale are
given in Fig. 1. The first chemical treatment combines MoO3 H2 MoO2 H2 O 1
MoO3 with sodium hydroxide to form a sodium molybdate Kennedy and Bevan [24] concluded that the reduction of
(Na2MoO4) precipitate. Sodium molybdate is commonly MoO3 followed the general mechanism proposed by
used as a fertilizer, but has also been used in laboratory Anderson [25] for the reduction of metal oxides:
settings to produce various pure molybdenum products.
The second chemical treatment synthesizes ammonium 1. Chemisorption of hydrogen on the surface
molybdates in purified forms. Ammonium heptamolybdate 2. Desorption of water, creating anion vacancies
((NH4)6Mo7O24) is commonly used in solution chemistry 3. Diffusion of the vacancies from the surface into the
because it dissolves readily in water, forming MoO42-(aq) bulk
species. The final purification method is sublimation and 4. Saturation of the crystal lattice with defects
condensation of MoO3. Further chemical processing of these 5. Nucleation of the new phase
products produce other common laboratory molybdenum
precursors such as molybdenum chlorides (MoCl2, MoCl4, Kennedy proposed that the rate-limiting step in
MoCl5), molybdic acid (MoO3H2O), molybdenum acetate reduction was the diffusion of anion vacancies (oxygen
(Mo2(O2CCH3)4), molybdenum ethoxide (C10H25O5Mo), vacancies, in this case) from the surface to the bulk,
various other molybdenum compounds, and molybdenum despite the reaction being linear with respect to time.
metal. Diffusion limited processes are not normally linear in
nature [26]. It was suggested that anisotropy of the
Direct reduction of MoO3 orthorhombic MoO3 crystal structure explained this unu-
sual phenomenon; specifically, that oxygen vacancy dif-
Traditionally, MoO2 has been produced by the reduction of fusion occurs primarily along {010} faces, and that the
MoO3 in dry hydrogen at temperatures between 450 and surface area would not decrease greatly as the reaction
proceeds. Several studies seem to confirm the importance
of oxygen transport on {010} MoO3 planes during
reduction [27, 28]. Burch [29] reinforced the reaction
mechanism proposed by Kennedy, but proposed that the
rate-limiting step for reduction of MoO3 is the chemi-
sorption of hydrogen on active surface sites. It is sug-
gested that this mechanism provides a more reasonable
explanation for the linear rate behavior.
Later, the importance of multiple mixed-valence oxides
during reduction was stressed, most notably the ortho-
rhombic Mo4O11 phase. Previous failure to detect mixed-
valence phases in some studies is attributed to limitations
of the characterization technique employed. Standard in
situ X-ray diffraction (XRD) techniques may not have the
time resolution or sample penetration necessary to accu-
rately and comprehensively analyze a sample in reduction.
The detection of Mo4O11 prompted a consecutive two-step
Fig. 1 Processing of technical molybdenum oxide reduction process [30].

123
J Mater Sci (2012) 47:20572071 2059

Step 1: reduction conversion (Eq. 7) of MoO3 to Mo4O11 with


4MoO3 H2 Mo4 O11 H2 O 2 hydrogen.
Partial reduction:
Step 2:
4MoO3 H2 Mo4 O11 H2 O 7
Mo4 O11 3H2 4MoO2 3H2 O 3
Lalick did not conclude whether comproportionation or
Lalik et al. [31] found both of the sequential reaction competitive nucleation is the actual reaction process
processes unlikely based on in situ neutron powder involved in MoO3 reduction.
diffraction (NPD) studies. This study used a 30/70 H2/He Detection limitations are likely not the only reason that
mixture flowing at 30 cm3/min for reduction at 550 C. The mixed valence phases are not found in many in situ
results are summarized in Fig. 2, where the concentration of reduction experiments. It has been proposed that the
the MoO2 and Mo4O11 phases increase at a similar rate appearance or lack of appearance of mixed valence phases
initially. The MoO2 concentration plateaus while the MoO3 is a function of the reaction conditions and starting material
is converted to the mixed valence phase. Finally, the mixed properties, especially oxygen vacancy concentration [29]
valence phase is fully converted to MoO2. and initial MoO3 particle size [32]. In addition, Mo4O11 was
Two reaction mechanisms were proposed to explain the observed upon reduction with wet hydrogen, while Mo2O5
compositional progression in this experiment. The first, was observed upon reduction with dry hydrogen [29].
a comproportionative reaction, explains the initial rise The importance of other mixed valence molybdenum
and plateau in MoO2 concentration. Comproportionation oxides has been emphasized more recently. Using in situ
defines a reaction in which a species present in two dif- XRD and X-ray absorbance spectroscopy (XAS), Ressler
ferent oxidation states, combines to form an intermediate et al. [33] reported the detection of Mo18O52 mixed valence
oxidation state, as seen in Eq. 5. structure during the reduction of MoO3 with propene.
Full reduction: Cariati et al. [34] also identified Mo5O14 and Mo14O47 in
MoO3 H2 MoO2 H2 O 4 addition to Mo4O11 and Mo18O52 by spectroscopic analysis.
Nanowires containing MoO3 and conductive Mo5O14 have
Comproportionation: recently been investigated in conductive composites [35]. A
MoO2 3MoO3 Mo4 O11 5 density functional study examined MoxOy (x = 13,
y = 19) clusters, and determined that MoO3, Mo?, MoO?,
Partial reduction:
MoO2?, and MoO3? are the building blocks of larger, more
Mo4 O11 3H2 4MoO2 3H2 O 6 complex mixed valence phases [36]. A variety of hydrogen
molybdenum bronzes (HxMoO3) have also been identified
The second mechanism, competitive nucleation, suggests
[37, 38] and are of interest because of unique physical and
that MoO2 is never a reactant in the formation of Mo4O11, but
electrical properties, such as the topotactic insertion of
rather that kinetic considerations cause the plateau in MoO2
hydrogen into MoO3 and highly anisotropic transport
and the preferential formation of Mo4O11 in the intermediate
properties [39]. These phases are generally not seen in the
stages. In this case, Eq. 5 is replaced by a direct partial
final products, but should not be discounted as unimportant
during molybdenum oxide reduction.
Schulmeyer [40] proposed chemical vapor transport
(CVT) for the reduction of MoO3. As illustrated in Fig. 3, a
gas transport phase (TP1) is formed from MoO3 (S1) and
subsequently nucleated as MoO2 (S2) over time (t ). This
transport process continues until only the S2 product
remains.
In order to account for the formation of Mo4O11, a
second transport phase (TP2) is assumed to have formed in
the process. Therefore, the overall reaction is given by
Chemical vapor transport:

MoO3 ! TP1 ! Mo4 O11 ! TP2 ! MoO2 8


Schulmeyer gives no mention of the nature or speciation of
either transport phase, but CVT is supported by the
Fig. 2 Fractional content of molybdenum oxide during hydrogen observed lack of similarity in particle size and morphology
reduction at 550 C. [data from Ref. 31] between reactant and product.

123
2060 J Mater Sci (2012) 47:20572071

Fig. 3 Schematic
representation of molybdenum TP1
oxide reduction by CVT
t t t

S1 Nucleus of S2 Shrinking core Growing Nucleus S2


of S1 of S2

Molybdenum dioxide prepared by CVT shows the agent in this method, however, it is unclear whether the
presence of several oxide phases coexisting in a multilayer reaction takes place at the solidliquid interface or fully in
structure resembling that of a coreshell arrangement. The solution. MoO3 is known to show moderate solubility in
presence of multiple oxide phases, particularly in a core water, forming hexavalent molybdate (MoO42-) or dimo-
shell arrangement, can be determined X-ray photoelectron lybdate (Mo2O72-) ions. A 3:1 volume mixture of deionized
spectroscopy (XPS). Table 1 gives the molybdenum water and glycol was found to produce the most pure
3d binding energies for the metal in its four most common monoclinic MoO2 phase. This product was nanocrystalline:
oxidation states. about 7 nm particle diameter according to X-ray peak
broadening. Higher ethylene glycol content was found to
Solution based approaches to MoO2 synthesis produce H0.93MoO3, a hydrogen molybdenum bronze, which
has attracted some interest as a hydrogen storage material.
The reduction of MoO3 to MoO2 with hydrogen is relatively Our group has used this method extensively to make catalytic
simple and commercially viable. However, other methods materials (shown for example in Fig. 4), which are active
have been investigated with objectives including MoO2 toward the partial oxidation of hydrocarbons.
particle size reduction, morphological alteration, and surface A study by Naouel et al. [42] clarifies the reaction
area maximization. Many of the studies creating unique process of reducing molybdenum trioxide to molybdenum
morphologies seek a high surface area for applications, such dioxide by hydrothermal synthesis. A suspension of MoO3,
as LIB anodes. High surface area provides a greater active para-phenylenediamine (NH2C6H4NH2), and water was
site density for lithium ion intercalation, and thus a higher charged to a sealed vessel for 24 h at temperatures ranging
charge density for the battery. High surface area is also from 160 to 220 C. The following reaction sequence is
beneficial for catalysis as more active sites per volume allow proposed.
for greater reactant flow rates and more efficient reforming. Hydrolysis:
Many solution phase synthesis methods use hydrother-
2MoO3 H2 O Mo2 O7 2 2H 9
mal processes to produce MoO2 directly. Because pro-
cesses at temperatures below 100 C have been largely Reduction:
unsuccessful in precipitating MoO2 in solution, sealed

10
Mo2O72 +22H+ + 2N
NH2C6H4NH2 = 2M
MoO
O2 + 3H
H2O + 2

autoclaves are frequently employed for hydrothermal pro- These reactions produce clusters of nanoparticles that
cesses. Chen et al. [41] charged a MoO3, ethylene glycol, are bound together with a layer of organic material into
water suspension to a Teflon-lined stainless steel auto- highly uniform microspheres, as shown in Fig. 5.
clave for 12 h at 180 C. Ethylene glycol is the reducing Wang et al. [43] investigated a template assisted
hydrothermal method with MoO3 nanofibers as templates.
Table 1 Binding energy of the most common oxidation states of Nanofibers were added to an autoclave with an acid
molybdenum
(pthalic, nitric, or sulfuric) and a reducing agent (ethanol or
Oxidation state Binding energy (eV) hydrazine). The products seem to retain the overall nano-
Mo3? 228.7
fiber morphology, but after reduction are highly porous and
Mo4? 229.5
composed of agglomerated nanoparticles.
Mo5? 231.5
Yang et al. [44] also used MoO3 as a precursor for
nanostructured MoO2 in a hydrothermal process. MoO3,
Mo6? 232.7
assisting agent Fe2O3, reducing agent ethylene diamene,

123
J Mater Sci (2012) 47:20572071 2061

different morphologies [46]. The product with ethylene


glycol was sheet like, with an average diameter around
250 nm and average thickness less than 100 nm. The eth-
ylene diamine product was bar-like with length up to
15 lm, breadth of approximately 400 nm, and thickness
approximately 100 nm.
Guo et al. [47] used AHM as a precursor for the syn-
thesis of a hollow microsphere morphology MoO2. AHM
was added to water with cetyltrimethyl-ammonium bro-
mide (CTAB) and ethanol under constant stirring. After the
solution became turbid, it was transferred to an autoclave,
sealed, and heated to 180 C for 5 days. The end product,
seen in Fig. 6b, shows porous, micron-scale, hollow
spheres. Crystallite size was found to be approximately
35 nm according to X-ray peak broadening. Only a very
Fig. 4 Scanning electron microscope (SEM) image of nanoparticle minor weight loss was observed by TG up to 335 C,
MoO2 synthesized by hydrothermal method
indicating that CTAB is probably not involved in the final
MoO2 product. For comparison, the authors compared
and water were charged to an autoclave for 5 days at CTAB (a cationic surfactant), with dioctadecyldimethy-
200 C. The product of this reaction was a highly porous, lammonium bromide (DCAB, another cationic surfactant),
nano-sheet, tremella-like microstructure, shown in PEO20-PPO70-PEO20 (P123, a non-ionic surfactant), PEG-
Fig. 6a. The sheet thickness is \10 nm. 400 (non-ionic), and sodium laurylsurfonate (SDS, anionic)
Several authors have investigated hydrothermal syn- in the same process. Only the MoO2 product made with
thesis of MoO2 using ammonium heptamolybdate (AHM) DCAB was hollow spherical. The authors suggest that
as the molybdenum precursor. AHM is advantageous in cationic surfactants have electrostatic interactions with
that it is safe, inexpensive, and highly soluble in water. negatively charged molybdate ions in solution and the
Liang et al. [45] dissolved AHM crystals in distilled water, structure defining surfactant favors the microsphere
then added hydrazine hydrate (N2H4H2O) and charged the morphology.
solution to an autoclave for 12 h at 130 C. In this case, the A novel method by Liu et al. [48] produced amorphous
reduction clearly proceeds in solution with hydrazine act- MoO2 nanoparticles in solution using c-ray radiation at
ing as the reducing agent. The product was amorphous, but ambient temperature and pressure. AHM was dissolved in
crystallized to pure monoclinic MoO2 upon heating in water with a surfactant (either polyvinyl alcohol or sodium
argon to 400 C. Thermogravimetric (TG) and differential dodecyl sulfate) and isopropanol (as a hydroxyl radical
thermal analysis (DTA) indicated the as-synthesized scavenger). After irradiation in 70 kCi 60Co c-rays, MoO2
material lost adsorbed water at about 170 C, and crystal- was precipitated. The product was held at 300 C for 12 h
lization occurred at 370 C. After heat treatment, SEM in argon to form a crystalline material. X-ray diffraction
reveals loosely agglomerated 400 nm spherical particles. In showed MoO2 with traces of MoO3. TEM found particles
similar experiments with ethylene glycol or ethylene dia- narrowly distributed between 8 and 30 nm in size.
mine added as surfactants, pure MoO2 was produced in Increasing concentration of MoO3 seemed to correspond to

Fig. 5 Schematic representation and corresponding transmission electron microscope (TEM) image of MoO2 microsphere morphology. (Image
reprinted with permission [42])

123
2062 J Mater Sci (2012) 47:20572071

Fig. 6 SEM images of MoO2


microstructures a Tremella-like
[44], b hollow microsphere
[47], c nanocast mesoporous
[9], d nanowire [56], e pinaster
[57], f nanostar [58] (Images
reprinted with permission)

a minor increase in particle size. Higher concentrations of Hydrolysis:


both surfactants corresponded to a decrease in particle size. BH 
4 2H2 O BO2 4H2 11
Although the mechanism of reduction is not discussed in
detail, it is suggested that c-ray radiation produces hydrate This study employed a sodium molybdate precursor and
electrons (e- aq), which are strong reducing agents. These in produced mixtures of MoO2 and Na0.88Mo6O17 bronze. The
turn reduce molybdates in solution to MoO2 nanoparticles. product is initially amorphous and is crystallized fully upon
Shi et al. [9] used nanocasting on KIT-6 silica templates heating to 350 C. The generalized competing reactions to
to produce ordered mesoporous MoO2 nanostructures. produce the reduced oxide and the bronze are given as
Phosphomolybdic acid (PMA) filled the silica mesopores, MoO2 precipitation:
and then was heated to 500 C in 10% H2/Ar. During this Na2 MoO4 NaBH4 2H2 O NaBO2 MoO2
heat treatment PMA is decomposed, reduced from Mo6? to 2NaOH 3H2 12
Mo4?, and crystallized as MoO2. The SiO2 template is then
removed using 4% HF. SEM and TEM confirm a highly Bronze precipitation:
ordered MoO2 structure. The Brunauer, Emmett, Teller Na2 MoO4 NaBH4 3  g  dH2 O
(BET) method estimates a surface area of 66 m2/g. A high NaBO2 Nag MoO3d 2  dNaOH 13
resolution SEM image of this structure is given in Fig. 6c.
4  0:5g  dH2
Tsang et al. [49] used alkali-metal borohydrides (e.g.,
NaBH4) to reduce transition metal oxides in solution at Greater borohydride concentration was found to
ambient temperature and pressure. The borohydride increase MoO2 yield. Higher borohydride concentrations
quickly hydrolizes in solution by were thought to strengthen the reducing power of the

123
J Mater Sci (2012) 47:20572071 2063

solution, thereby promoting the lower-valent MoO2. Pure water and adjusted to pH 8.5, producing MoO42- ions in
MoO2 is formed at pH 10 and 1, while mixtures of MoO2, solution. Then, with the HOPG plate employed as the
Mo4O11, and Na0.88Mo6O17 were produced at pH 4 and 7. working electrode in a three-electrode cell, MoO2 was
This phenomenon is explained by two opposing effects. deposited by
First, the reducing power of tetrahydroborate increases Electrodeposition:
with decreasing pH. Second, at lower pH molybdate ions
MoO4 2 2H2 O 2e MoO2 4OH 14
condense, making reduction more difficult. So, at high pH
monomeric molybdates, MoO42-, are easily reduced; and The nucleation energy barrier is lower on step edges than
at low pH the hydroborate has high reducing power. The on terraces of a clean surface of HOPG, therefore, nano-
detrimental (to MoO2 formation) combined effect at more wires can be grown preferentially at step edges, as seen in
neutral pH results in the formation of other species. Fig. 7. MoO2 nanowires can be grown specifically at step
In a similar study, the reduction of aqueous K2MoO4 edges between -0.6 and -1.10 V. Short deposition times
with aqueous KBH4 was investigated systematically [50]. produced nanowires with diameters of 1015 nm, while
The results were similar to the study with sodium boro- longer depositions produce diameters up to 700 nm.
hydride. An amorphous phase is formed, which crystallizes Patil et al. [55] studied thin films electrodeposited from
at 350400 C. However, unlike the previous study, the AHM solution onto fluorine-doped tin oxide (FTO) coated
competing effects of reducing power and condensation of glass. The resulting films are amorphous, and crystallize upon
molybdates in solution does not produce counterintuitive argon annealing to 450 C. Film thickness increased with
trends. In this study, reducing power of the borohydride deposition time up to a plateau at about 20 s. It is proposed that
increases with increasing concentration of borohydride at MoO42- ions are reduced to MoO2 with bonded hydroxide,
any given pH. Similarly, decreasing pH always results in a and that annealing eliminates the bonded hydroxide. Band gap
higher concentration of MoO2 in the product. This study measurements are reported to vary between 2.6 and 2.8 eV,
also produced Mo4O11 as a secondary phase, along with depending weakly on film thickness.
KxMoyOz bronzes.
Koziej et al. [51] used a non-aqueous solution method to Non-solution approaches to MoO2 synthesis
synthesize MoO2 nanoparticles. Molybdenum dioxide
chloride (MoO2Cl2) was charged with benzyl alcohol and Several experiments have attempted to make unique mor-
acetophenone to an autoclave for 48 h at 200 C. This phology MoO2 by non-solution approaches. These pro-
reaction produces 2 nm spheres, which aggregate into cesses are essentially thermal evaporation of molybdenum
larger (up to 200 nm in length), highly crystalline, elon- metal under vacuum. Zhou et al. [56] grew MoO2 nano-
gated nanoparticles. Interestingly, the authors reported a wires on a silicon substrate by heating a molybdenum boat
phase transition (from hexagonal to monoclinic) as the to 1100 C in flowing argon. The resulting nanowires,
nanospheres aggregated to form larger elongated particles. shown to be pure monoclinic MoO2, grew perpendicular to
Similar phenomena have been observed in numerous the substrate. The wires grow straight with diameter of
materials, including alumina [52] and zirconia [53]. 50120 nm, and are up to 4 lm long, as shown in Fig. 6d.
Zach et al. [54] used electrochemical step edge deco- The proposed mechanism is given in Fig. 8. Unstable
ration to deposit MoO2 nanowires on higly ordered pyro- MoOx condenses on the surface, and reacts with residual
lytic graphite (HOPG). Sodium molybdate was dissolved in oxygen in the chamber to form stable MoO2. After the seed

Fig. 7 Schematic
representation and SEM image
of edge-step electrodeposited
MoO2 nanowires (Image
reprinted with permision [54])

123
2064 J Mater Sci (2012) 47:20572071

of 4% oxygen. Molybdenum metal, MoO2, and MoO3 were


identified, and the particles ranged from 10 to 25 nm. This
material performed remarkably well in as a negative elec-
trode in a novel lithium ion battery (See Table 3).

MoO3 synthesis methods

Because MoO3 is so easily reduced by either hydrogen or


another reducing gas, it is pertinent to briefly cover the
Fig. 8 Mechanism of thermal evaporation for deposition of MoO3 products synthesized in research applications. Sev-
nanowires eral studies have produced unique MoO3 morphologies by
hydrothermal reactions, mimicking MoO2 synthesis meth-
ods, with the use of an autoclave at temperatures between
is planted on the substrate, MoO2 grows preferentially in 120 and 200 C. The products include nanoparticles
the wire direction. The authors also successfully formed [60, 61], nanorods [62, 63], nanobelts [6467], and nanof-
MoO3 nanowires by heating the as-grown MoO2 nanowires ibers [68]. Ambient pressure solution products include
to 400 C in an argonoxygen mixture. Alternatively, aerogel structures [69], whiskers [70], nanoparticles [71],
molybdenum metal nanowires were formed by heating as- nanorods [72], nanoplatelets [73], nanoflowers [2], and
grown MoO2 nanowires to 800 C in flowing hydrogen. lamellar mesostructured mixed molybdenum (V/VI) oxide
Zhang et al. [57] used thermal evaporation to deposit a [74]. Films of MoO3 have been formed by electrodeposition
unique pinaster-like structured MoO2 film on a silicon sub- [75], RF sputtering [7, 76, 77], and vacuum evaporation [5].
strate. The substrate was placed over a boat filled with Figure 9 gives some examples of the range of MoO3 mor-
molybdenum metal powder, and was heated to 1300 C phologies recently synthesized.
under vacuum. X-ray diffraction showed a pure monoclinic Any of these hexavalent oxide products can easily be
MoO2 phase. The unique pinaster morphology has a trunk converted to MoO2 by a simple hydrogen reduction above
with a diameter of 0.83 lm and a length about 12 lm. 450 C. However, retaining MoO3 morphology, particle
Approximately 8 nm nanowires protrude from the trunk size, or surface area is challenging. If the reduction of
along its length. Also, a relatively long (12 lm) nanowire MoO2 to MoO3 is fully in solid state, it involves a signif-
grew at the end of each trunk, which the authors identified as icant density change as well as a change in crystal struc-
a potential field emission point. Figure 6e shows an SEM ture. If the reduction involves vapor transport, there is little
image of this unique morphology. Similar to Zhou et al., the reason to expect the condensed MoO2 product to resemble
authors propose a vaporsolid (VS) mechanism for growth its MoO3 precursor. Studies have suggested overall particle
with residual chamber oxygen reacting with unstable gas- growth upon reduction from molybdenum oxide [78, 79].
eous molybdenum to form MoO2. Alternatively, Hu et al. synthesized MoO2 nanorods by
In a similar study, Khademi et al. [58] grew nanostar cleaving MoO3 nanobelts during reduction in hydrogen at
MoO2. A boat of molybdenum metal powder was heated to 550 C. The proposed growth mechanism includes break-
1300 C and molybdenum oxide condensed on a silicon ing bonds selectively along the [001] direction with crack
substrate. Star-shaped nanorods grew perpendicular to the initiation at defects, such as oxygen vacancies [80]. This
surface with a layer of spherical nanoparticles separating mechanism might serve to explain the result of Yang et al.
the substrate and the star-shaped nanorods. XRD showed [81], where MoO3 reduced with ethanol vapor retained the
crystalline Mo4O11 and crystalline MoO2, and TEM indi- MoO3 bulk microstructure while cleaving into nanocrys-
cated that the nanorods were primarily MoO2, while the tals. Burch [29] also finds MoO2 to be a solid-state process
spherical layer was Mo4O11. XPS showed a mixture of where MoO2 is likely to retain the general morphological
Mo6?, Mo5?, Mo4?, and Mod? (0 \ d \ 4). The wall habit of the MoO3 from which it was reduced.
thickness of the MoO2 nanostars (shown in Fig. 6f) is
approximately 50 nm. Nanoparticle molybdenum oxide
Dillon et al. [59] use hot-wire chemical vapor deposition
(HWCVD) to synthesize molybdenum nanoparticles for Nanomaterials have been of great interest for decades
application in rechargeable LIB anodes. In this procedure, a [8284], and are especially applicable to catalysis [85, 86],
molybdenum filament is resistively heated to *1400 C in fuel cells [87], and lithium ion batteries [88, 89]. Table 2
an environment of flowing argon and oxygen. The product summarizes molybdenum oxide synthesis methods,
collects on the quartz tube as the filament slowly oxidized. including the average smallest dimension as stated in the
Optimized synthesis occurred at 300 C with an atmosphere reference.

123
J Mater Sci (2012) 47:20572071 2065

Fig. 9 SEM images of MoO3 microstructures. a Nanobar [62], b nanobelt [65], c nanopalatelet [73], and d nanoflower [2] (Images reprinted
with permission)

Table 2 Summary of nanoscale


Synthesis method Product Size (nm) Morphology Reference
molybdenum oxide synthesis
techniques Hydrogen reduction of nanobars MoO2 100200 Nanorod 80
Ethanol vapor reduction MoO2 100 Aggregated sphere 81
Hydrothermal MoO2 50200 Sphere, bar, flake 45, 46
Hydrothermal MoO2 2050 Sphere 41
Hydrothermal MoO2 80200 Nanofiber, particle 43
Hydrothermal MoO2 10 Tremella 44
Hydrothermal MoO3 6090 Nanorod 62
Hydrothermal MoO3 50900 Nanobelt 64
Hydrothermal MoO3 50 Nanorod 63
Hydrothermal MoO3 50400 Nanobelt 66
Hydrothermal MoO3 50100 Nanobelt 61
Hydrothermal MoO3 20100 Nanofiber 65
Hydrothermal MoO3 50150 Nanorod 72
Hydrothermal MoO3 200300 Nanobelt 67
Microwave hydrothermal MoO3 50500 Nanoflower 2
Solvothermal MoO2 520 Nanorod 51
Solvothermal MoO3 100 Sphere 60
Solvothermal MoO3 100 Fiber 68
Solgel MoO3 10100 Sphere 71
Solgel MoO3 10100 Nanoplatelet 73
Gamma-ray radiation MoO2 830 Sphere 48
Nanocast on silica tempalte MoO2 8 Mesoporous 9
Hot-wire CVD MoO2 530 Sphere 59
Thermal evaporation MoO2 50100 Nanowire 56
Thermal evaporation MoO2 50 Nanostars 58
Thermal evaporation MoO2 8 Pinaster-like 57
Electrochemical deposition MoO2 1015 Nanowire 54
RF sputtering MoO3 10100 Thin film 76

123
2066 J Mater Sci (2012) 47:20572071

Applications Catalytic testing has shown success in reforming iso-


octane, premium gasoline, Jet-A, and biodiesel with high
Partial oxidation of hydrocarbons hydrogen yield. A significant issue in reforming commer-
cial fuels is the high concentrations of sulfur and coking
Molybdenum oxides have been used in various hydrocar- precursors. Sulfur poisoning and coking quickly deacti-
bon reforming applications for decades. Griffith [9092] vates traditional catalysts, such as nickel. Molybdenum
was the first to evaluate MoO2 as a catalyst for the dioxide has demonstrated excellent sulfur tolerance and
decomposition of hydrocarbons. MoO3 has been utilized coking resistance, as shown in Fig. 11a, b, respectively.
for the oxidation of propene [93, 94]. Partially reduced The reforming environment is critical for the effective
MoO3 and MoO2 have been investigated for the isomeri- performance of the MoO2 catalyst. Exhaustive thermody-
zation of alkanes [13, 95, 96], hexanes [97], and heptanes namic calculation and experimental study [14] has shown a
as well as the dehydration of propanol [10]. relatively narrow stability window for MoO2 in oxygen-
A bifunctional mechanism has been proposed to explain containing reforming environments at 550900 C.
the catalytic activity of MoO2 for the isomerization of Figure 12 shows calculated stability of the species in the
hydrocarbons [98]. First, metallic sites (Mo4? in MoO2) MoOC system at 850 C. At this temperature, oxygen to
dissociate hydrogen to produce active hydrogen atoms. carbon (O2/C) ratios above 1.0 promote oxidation to MoO3,
These active hydrogen atoms bond with surface oxygen to while O2/C ratios below 0.5 may promote coke formation
form Brnsted acid functional groups [12]. There is or reduction to molybdenum carbide (Mo2C).
ongoing work focusing on MoO2 as a catalyst for the MoO2 catalysts are severely limited by this environ-
partial oxidation of hydrocarbons, seeking the efficient mental instability. A proposed solution involves doping
production of hydrogen [14, 99, 100]. Using isooctane as a MoO2 with redox stable species such as Ti, Si, and Sn.
model fuel, partial oxidation can be given by Altering bond energies in the crystal lattice should change
Partial oxidation: the redox properties of the catalyst, while retaining the
C8 H18 4O2 , 8CO 9H2 15 catalytic activity. Studies on silicon [102] and rare earth
oxide [103] doping have already shown potential to
The proposed mechanism for this catalytic reaction is similar increase stability. Based on past success [e.g., 104, 105],
to that described by Mars and Van Krevelen [101], whereby research into titanium doping is ongoing, and seeks to
the catalyst undergoes cycles of oxidation and reduction. evaluate the phase stability and catalytic activity of a single
First, the hydrocarbon is decomposed on the oxide surface phase MoxTi(1-x)O2 catalyst.
and Mo4? ions gain electrons from hydrocarbon molecules.
Next, lattice oxygen diffuses to the surface of the catalyst and Direct solid oxide fuel cell (SOFC) anode material
combines with smaller hydrocarbon molecules, forming
carbon oxides, hydrogen, and water. Finally, gaseous oxygen Solid oxide fuel cells (SOFC) are an attractive option in
replenishes the lattice oxygen lost in the reaction. This alternative energy production. Traditional SOFC anodes
mechanism is summarized in Fig. 10. are a nickelzirconia cermet where hydrogen reacts with

Fig. 10 Illustration of the


MarsVan Krevelen mechanism

123
J Mater Sci (2012) 47:20572071 2067

(a) (b)
100%
Hydrogen yield, Carbon conversion 90%
80%
70%
60%
50%
40%
30% Carbon conversion: 0 ppmw BT
Carbon conversion: 1000 ppmw BT
20%
Hydrogen yield: 0 ppmw BT
10%
Hydrogen yield: 1000 ppmw BT
0%
0 50 100 150 200 250 300 350 400 450
Time on-stream (minutes)

Fig. 11 a Catalytic performance of MoO2 at different benzothiophene (BT) concentrations and b comparison between the performance of MoO2
and a nickel catalyst for the partial oxidation of premium gasoline

Fig. 12 Ternary phase diagram O


MoOC at 850 C and 1 atm 1: MoO3(L) + gas Standard
2: MoO2+MoO3(L)+gas Gibbs Free Energy (J/mol)

3: Mo+MO2C+MoO2 -2 x 104
4: Mo2C+MoO2+gas 0.1 0.9
5: MoC+Mo2C+gas
6: MoC+gas 1 -12x 104
0.2 0.8
7: MoC+C+gas

0.3 2 0.7 (O2/C=1.17)


Pure MoO2

0.4 0.6 (O2/C=0.75)

0.5 4 6 0.5 (O2/C=0.50)

0.6 0.4

0.7 0.3

0.8
3 5 7 0.2

0.9 0.1

Mo C
0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9

charge-carrying oxygen ions to form water. Nickel pro- increase the applicability of SOFC power generation
vides the electrical conductivity necessary for charge immensely. Not only does eliminating a reformer save
transfer. Zirconia ensures a good interface with the elec- space, weight, and complexity, but also using commercial
trolyte and continues oxygen ion conductivity into the fuel allows power generation anywhere there is gasoline,
porous anode. Traditional SOFC anodes require pure H2 as Jet-A, or any other available hydrocarbon fuel.
fuel, and are subject to deactivation because of coke for- The addition of small amounts (about 1%) molybdenum
mation [106] or sulfur poisoning [107] if fueled by com- to a traditional Ni/yttria-stabilized zirconia (YSZ) anode
mercial fuels. Because of its fuel flexibility and electrical was found to vastly improve coking tolerance in fuel cells
conductivity, MoO2 has been identified as a promising running on methane [108]. Ioroi et al. [109] tested carbon-
direct-fueled SOFC anode material [16, 17]. Eliminating a supported Pt/MoOx composite anodes in proton exchange
pre-reformer and enabling use of commercial fuels would membrane (PEM) fuel cells, and found them to be highly

123
2068 J Mater Sci (2012) 47:20572071

CO tolerant at 100 ppm in H2. Other studies have found nanotubes, modified graphite) x has been shown to exceed
improved CO tolerance when molybdenum was added to 1, and storage capacity has exceeded the theoretical value
PtRu catalysts for PEM fuel cells running on methanol [119]. Non-carbon materials, including MoO2, have also
[110] and hydrogen [111]. Recently, a highly conductive been extensively studied as LIB anodes. In MoO2 based
strontium molybdate (SrMoO4)/YSZ anode was tested anodes, intercalation occurs by
[112]. The molybdate was deposited on porous YSZ by wet Lithium intercalation in MoO2:
impregnation, and produced power when fueled by MoO2 xLi xe Lix MoO2 17
hydrogen and methane.
Research in our group is ongoing into fabricating and Murphy explored the possibility of lithium intercalation
testing pure MoO2 and MoO2 ? Ni/YSZ SOFC anodes. into MoO2 in 1978 [120]. Auborn and Barberio [121] first
Figure 13 shows a composite anode on a YSZ electrolyte. established the potential of a MoO2 anode in 1987, but also
Molybdenum oxide-based SOFC anodes have the same found several practical issues that needed addressing. First,
stability range as partial oxidation catalysts; therefore, there is a phase transition upon intercalation when the
similar doping procedures are under investigation to pro- concentration reaches Li0.5MoO2. Cycling was found to be
vide greater flexibility for the anode fuel environment. reversible above this composition (x [ 0.5), but was irre-
versible when discharged past x = 0.5. Once a cell is
Lithium-ion battery anodes discharged through the phase transition, the capacity
decreases rapidly with cycling. Second, particle size has a
As energy storage needs continue to increase, battery significant influence on charge capacity of the MoO2
technology must advance to allow greater efficiency. anode. Unmilled powder was found to reach as little as
Lithium ion batteries have been studied extensively since 10% of the theoretical charge capacity, whereas milled
the 1980s [113115] and have been very successfully material reached 95% of the theoretical value. This
commercialized. However, research continuously seeks observation suggests that ion diffusion through MoO2 is
improvements in electrode materials, especially in terms of limiting, and a radial concentration gradient is formed in
enhancing storage capacity [e.g., 88, 116, 117]. Activated large particles. With this in mind, much of the later
carbon has been the standard anode material for commer- research into MoO2 LIB anodes has examined high surface
cial LIBs and has a theoretical storage capacity of area materials.
372 mAh/g [118]. Intercalation, the insertion of Li? ions Manthiram and Tsang [122] studied amorphous MoO2?d
into the carbon anode, occurs by (d * 0.3) as a LIB anode. The results show a highly
Lithium intercalation in carbon: reversible capacity of *220 mAh/g between 3 and 1 V,
corresponding to a lithium insertion of *1.1. However,
xLi xe 6C Lix C6 16
cell cyclability decreased dramatically upon discharge
The theoretical storage capacity is reached if x = 1 upon below 1 V.
full intercalation. Using non-graphitic carbon (e.g., In a series of studies, Liang et al. investigated two dif-
ferent MoO2 morphologies as LIB anodes. First,

Fig. 14 LIB anode performance of MoO2 synthesized by a reduction


Fig. 13 SEM fracture surface image of a Ni/YSZ ? MoO2 anode on in ethanol vapor (times), and by a hydrothermal method (square,
a YSZ electrolyte circle, triangle). Cycling in the range 0.013.00 V versus Li metal

123
J Mater Sci (2012) 47:20572071 2069

Table 3 Summary of MoO2


Morphology Charge capacity (mAh/g) Cycles Reference
LIB anode performance
Amorphous particles 220 15 122
Particles (*300 nm) 400 40 45
Porous spheres 480 40 46
Aggregated nanoparticle 381 20 81
Nanoparticle 620 50 59
Tremella-like 538 20 44
Mesoporous 750 30 9
MoO2/C nanorods 760 20 123
MoO2/C nanowires 350 20 124
Elongated nanoparticles 300 200 51

aggregated 400 nm spherical particles were formed by a Ji et al. and Gao et al. [123, 124] investigated carbon/
hydrothermal method from AHM. Second, elongated 700 by MoO2 hybrid nanostructures as materials for LIB anodes.
100 nm particles were synthesized from an oxalate precursor Ji produced a 36.6 wt% anode that had a stable capacity
by a low temperature method. The initial charge capacities of 760 mAh/g between 3.0 and 0.0 V. Gaos MoO2/C
were found to be 769 and 484 mAh/g, respectively [45]. hybrid nanowires achieved a capacity of over 350 mAh/g
There was a rapid decrease in cyclability below 2 V. Upon at a rate of 1000 mA/g for 20 cycles. An initial capacity of
cycling, the capacity of the elongated grains increased 600 mAh/g was attained at a rate of 200 mA/g, but the
slightly until cycle 13, then decreased slowly until cycle 40, capacity decreased significantly over 20 cycles. Table 3
where the anode retained about 83% of the initial value [46]. summarizes the performance of MoO2-based anodes for
The authors claim the increase in capacity up to cycle 13 can lithium ion batteries.
be attributed to electrochemical grinding, whereby the Additional research is required in order to make MoO2 a
particle size in the anode decreases, exposing more surface viable LIB anode material, especially with regard to
area and thus improving capacity. charging rates and long term cycling stability.
Yang et al. [81] tested a MoO2 LIB anode made from
ethanol-reduced MoO3. TEM showed that the material was
composed of agglomerated nanoparticles with an average Conclusions
diameter of approximately 100 nm. The initial charge
capacity was 318 mAh/g at 5.0 mA/cm2, and no decrease Molybdenum dioxide exhibits growing potential for a
in capacity was observed over 20 cycles between 0.01 and number of applications. It can be synthesized in several
3.0 V, as seen in Fig. 14. About 85% of the reversible highly specific morphologies by a variety of techniques.
capacity was between 1.0 and 2.0 V. Another study by MoO2 is easily obtained by reduction of MoO3 by hydro-
Yang et al. [44] tested tremella-like nanostructured MoO2 gen or other reducing gases above 450 C. Studies also
(seen in Fig. 5a) as a LIB anode. The initial charge found success making a wide variety of MoO2 and MoO3
capacity was 538 mAh/g at 0.5 mA/cm2, but only 314 and products by solution-phase and gas-phase approaches.
259 mAh/g at 2.5 and 5 mA/cm2, respectively. In all three Several of these studies seek high surface area products for
cases, the capacity rose slightly up to about 8 cycles, and application in lithium ion battery anodes. Other groups
had approximately the initial value after 20 cycles. This have recently used MoO2 as a catalyst for partial oxidation
trend is seen in Fig. 14. In this study, a very high lithium logistics of fuels such as biodiesel, gasoline, and Jet-A. In
insertion value of up to 3.2 was achieved. an extension of this research, MoO2 is being tested as an
Shi et al. used nanocast mesoporous MoO2 (Fig. 5c) as anode material for direct-fueled solid oxide fuel cell anode
an LIB anode material [9]. The anode held a charge of that is resistant to coking and sulfur poisoning.
750 mAh/g with no decrease in capacity after 30 cycles.
This high capacity was attributed to the mesoporous Acknowledgements This work was supported financially by the
National Science Foundation GOALI Program (Grant No. CBET-
materials high surface area (66 m2/g) and low electrical 1034308) and by Boeing Commercial Airplanes.
resistivity (*0.01 Xcm).
The elongated nanoparticle structures investigated by
Koziej et al. [51] showed an initial charge capacity of References
375 mAh/g between 3.0 and 0.0 V. The LIB anode
shows very good cyclability, retaining 300 mAh/g after 1. Rabalais JW, Colton RJ, Guzman AM (1974) Chem Phys Lett
200 cycles. 29:131

123
2070 J Mater Sci (2012) 47:20572071

2. Wei G, Qin W, Zhang D, Wang G, Kim R, Zheng D, Wang L 43. Wang S, Zhang Y, Wang W, Li G, Ma X, Li X, Zhang Z, Qian
(2009) J Alloys Compd 481:417 Y (2006) J Cryst Growth 290:96
3. Yang YA, Cao YW, Loo BH, Ya JN (1998) J Phys Chem B 44. Yang LC, Gao QS, Zhang YH, Tang Y, Wu YP (2008) Elect-
102:9392 rochem Commun 10:118
4. Hussain OM, Rao KS, Madhuri KV, Uthanna S, Julien C (2003) 45. Liang Y, Yang S, Yi Z, Sun J, Zhou Y (2005) Mater Chem Phys
Proc Electrochem Soc 17:56 93:395
5. Yao JN, Hashimoto K, Fujishima A (1992) Nature 355:624 46. Liang Y, Yi Z, Yang S, Zhou L, Sun J, Zhou Y (2006) Solid
6. Ferroni M, Guidi V, Martinelli G, Sacerdoti M, Nelli P, State Ionics 177:501
Sberveglieri G (1998) Sens Actuator B 48:285 47. Guo CH, Zhang GJ, Shen ZR, Sun PC, Yuan ZY, Jin QH, Li
7. Mutschall D, Holzner K, Obermeier E (1996) Sens Actuator B BH, Ding DT, Chen TH (2006) Chin J Chem Phys 19:543
36:320 48. Liu Y, Qian Y, Zhang M, Chen Z, Wang C (1996) Mater Res
8. Sunu SS, Prabhu E, Jayaraman V, Gnanasekar KI, Gnanaseka- Bull 31:1029
ran T (2003) Sens Actuator B 94:189 49. Tsang CF, Manthiram A (1997) J Mater Chem 7:1003
9. Shi Y, Guo B, Corr SA, Shi Q, Hu YS, Heier KR, Chen L, 50. Tsang CF, Dananjay A, Kim J, Manthiram A (1996) Inorg Chem
Seshadri R, Stucky GD (2009) Nano Lett 9:4215 35:504
10. Benadda A, Katrib A, Barama A (2003) Appl Catal A Gen 251:93 51. Koziej D, Rossell MD, Ludi B, Hintennach A, Novak P,
11. Haber J, Lalik E (1997) Catal Today 33:119 Grunwaldt JD, Niederberger M (2011) Small 7:377
12. Al-Kandari H, Al-Khorafi F, Belatel H, Katrib B (2004) Catal 52. McHale JM, Auroux A, Perrotta AJ, Navrotsky A (1997) Sci-
Commun 5:225 ence 277:788
13. Ono Y (2003) Catal Today 81:3 53. Garvie RC (1978) J Phys Chem 82:218
14. Marin-Flores O, Turba T, Breit J, Norton MG, Ha S (2010) Appl 54. Zach MP, Inazu K, Ng KH, Hemminger JC, Penner RM (2002)
Catal A Gen 381:18 Chem Mater 14:3206
15. Marin-Flores O, Turba T, Ellefson C, Scudiero L, Breit J, Ha 55. Patil RS, Uplane MD, Patil PS (2006) App Surf Sci 252:8050
SY, Norton MG (2011) Appl Catal B. doi:10.1016/j.apcatb. 56. Zhou J, Xu NS, Deng SZ, Chen J, She JC, Wang ZL (2003) Adv
2011.03.035 Mater 15:1835
16. Marin-Flores O, Turba T, Ellefson C, Wang K, Breit J, Ahn J, 57. Zhang JY, Lui YG, Shi SL, Wang YG, Wang TH (2008) Chin
Norton MG, Ha S (2010) Appl Catal B Env 98:186 Phys B 17:4333
17. Marin-Flores O, Turba T, Ellefson C, Scudiero L, Breit J, 58. Khademi A, Azimirad R, Zavarian AA, Moshfegh AZ (2009) J
Norton MG, Ha S (2010) J Nanoelectron Optoelectron 5:1 Phys Chem C 113:19298
18. Gupta CK (1992) Extractive metallurgy of molybdenum. CRC 59. Dillon AC, Mahan AH, Deshpande R, Parilla PA, Jones KM,
Press, Boca Raton Lee SH (2008) Thin Solid Films 516:794
19. Hagg G, Magneli A (1944) Arkiv Kemi Min Geol 19 60. Kim WS, Kim HC, Hong SH (2009) J Nanopart Res. doi:
20. Magneli A, Andersson G (1955) Acta Chem Scand 9:1378 10.1007/s11051-009-9751-6
21. Marinder BO, Magneli A (1957) Acta Chem Scand 11:1635 61. Xia T, Li Q, Liu X, Meng J, Cao X (2006) J Phys Chem B
22. Kihlborg L (1959) Acta Chem Scand 13:954 110:2006
23. Kihlborg L (1969) Acta Chem Scand 23:1834 62. Fang L, Shu Y, Wang A, Zhang T (2007) J Phys Chem C 111:
24. Kennedy MJ, Bevan SC (1974) J Less Common Met 36:23 2401
25. Anderson JS (1948) Discuss Farad Soc 4:163 63. Lou XW, Zeng HC (2002) Chem Mater 14:4781
26. House JE (2007) Principles of chemical kinetics. Academic 64. Wang S, Zhang Y, Ma X, Wang W, Li X, Zhang Z, Qian Y
Press, Burlington (2005) Solid State Commun 136:283
27. Dufour LC, Bertrand O, Floquet N (1984) Surf Sci 147:396 65. Song RQ, Xu AW, Deng B, Fang YP (2005) J Phys Chem B
28. Smith RL, Rohrer GS (1996) J Catal 136:12 109:22758
29. Burch R (1978) J Chem Soc Farad Trans 74:2982 66. Li XL, Liu JF, Li YD (2002) Appl Phys Lett 81:4832
30. Sloczynski J (1995) J Solid State Chem 118:84 67. Chen Y, Lu C, Xu L, Ma Y, Hou W, Zhu JJ (2010) Cryst Eng
31. Lalik E, David WIF, Barnes P, Turner JFC (2001) J Phys Chem Commun 12:3740
B 105:9153 68. Patzke GR, Michailovski A, Krumeich F, Nesper R, Grunwaldt
32. Leisegang T, Levin AA, Walter J, Meyer DC (2005) Cryst Res JD, Baiker A (2004) Chem Mater 16:1126
Technol 1/2:95 69. Harreld JH, Dong W, Dunn B (1998) Mater Res Bull 33:561
33. Ressler T, Wienold J, Jentoft RE, Neisius T (2002) J Catal 70. Krishnan CV, Chen J, Burger C, Chu B (2006) J Phys Chem B
210:67 110:20182
34. Cariati F, Bart JCJ, Sgamellotti A (1981) Inorg Chem Acta 48:97 71. Ganguly A, George R (2007) Bull Mater Sci 30:183
35. Domenici V, Conradi M, Remskar M, Virsek M, Zupancic B, 72. Taurino AM, Forleo A, Francioso L, Siciliano P (2006) App
Mrzel A, Chambers M, Zalar B (2011) J Mater Sci 46:3639. doi: Phys Lett 88:152111
10.1007/s10853-011-5280-7 73. Wang G, Ji Y, Zhang L, Zhu Y, Gouma PI, Dudley M (2007)
36. Oliveira JA, De Almeida WB, Duarte HA (2003) Chem Phys Chem Mater 19:979
Lett 372:650 74. Yu X, Xu Z, Han S, Che H, Yan X, Liu A (2009) Colloid Surf A
37. Schollhorn R, Kuhlmann R, Besenhard JO (1976) Mater Res 333:194
Bull 11:83 75. Nagirnyi VM, Apostolova RD, Shembel EM (2006) Appl
38. Dickens PG, Birtill JJ, Wright CC (1979) J Sol State Chem Electrochem Corr Prot Met 79:1454
28:185 76. Oros C, Wisitsoraat A, Phokharatkul D, Limsuwan P, Tuantr-
39. Greenblatt M (1988) Chem Rev 88:31 anont A (2011) Adv Mater Res 213:98
40. Schulmeyer WV, Ortner HM (2002) Int J Ref Met Hard Mater 77. Comini E, Faglia G, Sberveglieri G, Cantalini C, Passacantando
20:261 M, Santucci S, Li Y, Wlodarski W, Qu W (2000) Sens Actuator
41. Chen X, Zhang Z, Li X, Shi C, Li X (2006) Chem Phys Lett B 68:168
418:105 78. Wang X, Liu J, Zhuang F, Xhao H, Li J (2010) Met Mater Trans
42. Naouel R, Touati F, Gharbi N (2010) Sol State Sci 12:1098 B 41:1067

123
J Mater Sci (2012) 47:20572071 2071

79. Radchenko PY, Panichkina VV, Radchenko OG (1999) Powder 102. Liang Y, Tracy C, Weisbrod E, Fejes P, Theodore ND (2006)
Mettal Met Ceram 38:429 Appl Phys Lett 88:081901
80. Hu B, Mai L, Chen W, Yang F (2009) ACS Nano 2:478 103. Wang J, Ren Z, Liu W, Zhou M (2009) Int J Ref Met Hard
81. Yang LC, Gao QS, Tang Y, Wu YP, Holze R (2008) J Power Mater 27:155
Source 179:357 104. Jacob KT, Shekhar C, Waseda Y (2008) J Am Ceram Soc 91:563
82. Krostoff RN, Koytcheff RG, Lau CGY (2007) Curr Sci 92:1492 105. Stengl V, Bakardjieva S (2010) J Phys Chem 114:19308
83. Roduner E (2006) Chem Soc Rev 35:583 106. Timmerman H, Sawady W, Campbell D, Weber A, Reimert R,
84. Cushing BL, Kolesnichenko VL, OConnor CJ (2004) Chem Ivers-Tiffee E (2008) J Electrochem Soc 155:B356
Rev 104:3893 107. Matsuzaki Y, Yasuda I (2000) Solid State Ionics 132:261
85. Zecchina A, Groppo E, Bordiga S (2007) Chem Eur J 13:2440 108. Finnerty CM, Coe NJ, Cunningham RH, Ormerod RM (1998)
86. Grunes J, Zhu J, Somorjai GA (2003) Chem Commun 18:2257 Catal Today 46:137
87. Arico AS, Baglio V, Antonucci V (2009) Nanotechnology for 109. Ioroi T, Yasuda K, Siroma Z, Fujiwara N, Miyazaki (2003)
the energy challenge. Wiley, Weinheim J Electrochem Soc 150:A1225
88. Bruce PG, Scrosati B, Tarascon JM (2008) Andew Chem Int Ed 110. Lima A, Coutenceau C, Leger JM, Lamy C (2001) J Appl
47:2930 Electrochem 31:379
89. Liu HK, Wang GX, Guo Z, Wang J, Konstantinov K (2007) J 111. Papageorgopoulos DC, Keijzer M, de Bruijn FA (2002) Electro
New Mater Electrochem Sys 10:101 Acta 48:197
90. Griffith RH, Plant JHG (1935) Proc R Soc Lond 148:191 112. Smith BH, Gross MD (2011) Electrochem Sol-State Lett 14:B1
91. Griffith RH, Lindars PR (1950) Nature 165:486 113. Scrosati B (2000) Electrochem Acta 45:2461
92. Griffith RH, Chapman PR, Lindars PR (1950) Disc Farad Soc 114. Daniel C (2008) J Mater 60:43
8:258 115. Megahed S, Ebner W (1995) J Power Sour 54:155
93. Peacock JM, Parker AJ, Ashmore PG, Hockey JA (1969) J Catal 116. Tarascon JM, Armand M (2001) Nature 414:359
15:373 117. Scrosati B, Garche J (2010) J Power Sour 195:2419
94. Peacock JM, Sharp MJ, Parker AJ, Ashmore PG, Hockey JA 118. Yamamoto O, Imanishi N, Takeda Y, Kashiwagi H (1995)
(1969) J Catal 15:379 J Power Sour 54:72
95. Galadima A, Anderson JA, Wells RPK (2009) Sci World J 4:15 119. Wu YP, Rahm E, Holze R (2003) J Power Sour 114:228
96. Katrib A, Logie V, Saurel N, Wehrer P, Hilaire L, Maire G 120. Murphy DW, Di Salvo FJ, Carides JN, Waszczak (1978) Mater
(1997) Surf Sci 377:754 Res Bull 13:1395
97. Tsigdinos GA, Swanson WW (1978) Ind Eng Chem Prod Res 121. Auborn JJ, Barberio YL (1987) J Electrochem Soc Electrochem
Dev 17:208 134:638
98. Katrib A, Mey D, Marie G (2001) Catal Today 65:179 122. Manthiram A, Tsang C (1996) J Electrochem Soc 143:L143
99. Marin Flores OG, Ha S (2009) App Catal A 352:124 123. Ji X, Herle S, Rho Y, Nazar LF (2007) Chem Mater 19:374
100. Marin-Flores O, Scudiero L, Ha S (2009) Surf Sci 603:2327 124. Gao Q, Yang L, Lu X, Mao J, Zhang Y, Wu Y, Tang Y (2010)
101. Vannice MA (2007) Catal Today 123:18 J Mater Chem 20:2807

123
Copyright of Journal of Materials Science is the property of Springer Science & Business Media B.V. and its
content may not be copied or emailed to multiple sites or posted to a listserv without the copyright holder's
express written permission. However, users may print, download, or email articles for individual use.

Вам также может понравиться