Você está na página 1de 42

Catalysis Reviews

Science and Engineering

ISSN: 0161-4940 (Print) 1520-5703 (Online) Journal homepage: http://www.tandfonline.com/loi/lctr20

Catalytic Properties of Single Layers of Transition


Metal Sulfide Catalytic Materials
Russell R. Chianelli , Mohammad H. Siadati , Myriam Perez De la Rosa , Gilles
Berhault , Jess P. Wilcoxon , Roby Bearden Jr. & Billie L. Abrams
To cite this article: Russell R. Chianelli , Mohammad H. Siadati , Myriam Perez De la Rosa ,
Gilles Berhault , Jess P. Wilcoxon , Roby Bearden Jr. & Billie L. Abrams (2006) Catalytic Properties
of Single Layers of Transition Metal Sulfide Catalytic Materials, Catalysis Reviews, 48:1, 1-41,
DOI: 10.1080/01614940500439776
To link to this article: http://dx.doi.org/10.1080/01614940500439776

Published online: 03 Feb 2007.

Submit your article to this journal

Article views: 892

View related articles

Citing articles: 112 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=lctr20
Download by: [Instituto Mexicano del Petroleo]

Date: 02 May 2016, At: 11:31

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Catalysis Reviews, 48:141, 2006


Copyright # Taylor & Francis Group, LLC
ISSN: 0161-4940 print 1520-5703 online
DOI: 10.1080/01614940500439776

Catalytic Properties of Single


Layers of Transition Metal
Sulfide Catalytic Materials
Russell R. Chianelli, Mohammad H. Siadati,
and Myriam Perez De la Rosa
Materials Research and Technology Institute, The University of Texas at El Paso,
El Paso, Texas, USA

Gilles Berhault
Institut, de Recherches sur la Catalyse, Villeurbanne, France

Jess P. Wilcoxon
Sandia National Laboratories, Albuquerque, New Mexico, USA

Roby Bearden, Jr.


Retired from ExxonMobil Process Research Laboratories, Baton Rouge, Louisiana, USA

Billie L. Abrams
Sandia National Laboratories, Albuquerque, New Mexico, USA
Single layer transition metal sulfides (SLTMS) such as MoS2, WS2, and ReS2, play an
important role in catalytic processes such as the hydrofining of petroleum streams,
and are involved in at least two of the slurry-catalyst hydroconversion processes that
have been proposed for upgrading heavy petroleum feed and other sources of hydrocarbon fuels such as coal and shale oils. Additional promising catalytic applications of
the SLTMS are on the horizon. The physical, chemical, and catalytic properties of
these materials are reviewed in this report. Also discussed are areas for future
research that promise to lead to advanced applications of the SLTMS.
Keywords

SLTMS, Catalytic processes, Transition metal sulfides

Received September 1, 2005; accepted October 23, 2005.


Address correspondence to Russell R. Chianelli, Materials Research and Technology
Institute, The University of Texas at El Paso, El Paso, TX, USA. E-mail: chianell@
utep.edu

R. R. Chianelli et al.

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

1.

INTRODUCTION

Catalysts based on the layered transition metal sulfides (LTMS) have played a
central role in petroleum refining since before WW II. The LTMS catalysts were
developed and applied precisely because of their ability to catalyze hydrocarbon-upgrading reactions such as hydrodesulfurization (HDS) and hydrogenation (HYD) in the presence of large amounts of sulfur and other
contaminants including nitrogen and metals. It is for this reason that catalysts
containing the transition metal sulfides MoS2 and WS2 operate in almost every
refinery in the world today to upgrade fuels by removing contaminants, as well
as, by increasing the hydrogen content of the feedstock through hydrogenation
of aromatic molecules. Moreover, many other important petroleum refinery or
chemical manufacturing processes have used these materials (reforming) in
the past or may use them in the future (alcohol) production. These applications
of LTMS catalytic materials were thoroughly reviewed in the 1973 work by
Weisser and Landa (1). The applications of LTMS catalytic materials were
recently updated (2 4).
Recent literature, to be discussed later, indicates that when the LTMS is
operated in a catalytic environment, particularly in hydrogen, it stabilizes
as single layers. It is the properties and application of these single layers of
transition metal sulfides, the SLTMS that are the subject of this review. It
is a major conclusion of this review that the SLTMS will continue to play a
major role in hydrocarbon upgrading and chemical processing in the future.
This conclusion is based on an analysis of the chemical and physical properties
of the SLTMS and the fact that much work has led to a deeper understanding
of their catalytic structure/function relation. This deeper understanding also
leads to an increased ability to synthesize and optimize their catalytic performance for application in specific catalytic processes. Improving catalytic
processes that include HDS, HYD, and HDN, as well as synthesis reactions
such as CO/H2, are crucial for the continuing economic prosperity of the
world economy. Furthermore, photocatalytic applications of the SLTMS are
also promising areas of research in several environmental and sustainable
energy applications.

2.

STRUCTURE OF THE LTMS

Catalysts based on MoS2 are the most commonly used LTMS catalysts in petroleum refining. The WS2 is used to a lesser extent but both have an identical
structure based on trigonal prisms of sulfur coordinated to Mo or W. The
trigonal prisms are strongly bonded in two dimensions that form the S Mo S
(S W S) sandwiches that can then stack to form three-dimensional crystals
of varying stacking arrangements. The single layer structure is indicated in
Fig. 1. It should be noted, however, that reduction of the single layers of

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Catalytic Properties of SLTMS

Figure 1: (A) Single layer structure of MoS2 and WS2. The yellow spheres are sulfur and the
purple spheres are Mo or W. (B) Single layer structure of ReS2. The yellow spheres are sulfur and
the green spheres are Re. Notice the Re Re bonds.

MoS2 could result in a distorted octahedral configuration in which the Mo IV is


reduced to Mo III. The Mo III then bonds with another Mo III resulting in a distortion of the closely packed sulfur layers. This distortion is usually described
as the distorted octahedral with respect to the Mo coordination. This form is
isostructural with layers of ReS2 (Fig. 1b) with Re IV being isoelectronic with
Mo III (5).

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

R. R. Chianelli et al.

Stacking of single layers of MoS2 occurs in several polytypes that reflect


different repeat units in the third dimension. The MoS2 is usually found in
the 2H (hexagonal) form indicating that the unit cell repeats along the hexagonal c-axis after two layers (0.612 nm perpendicular to the layers). Within each
layer the closely packed sulfur atoms determine the hexagonal a-axis of
0.312 nm. The 2H designation indicates that the second layer is translated to
one half with respect to the first layer. However, MoS2 can also be found in
the 3R (rhombohedral) form (6). The 3R designation indicates that the MoS2
layers repeat after three layers with the layers translated by one third with
respect to each other. Since this report is focused on single layers (SLTMS),
the stacking polytypes of MoS2 are of secondary importance. The reader is
directed for further information to an extensive review of LTMS structure
and properties by Wilson and Yoffe (7).

3.

FOLDING OF SINGLE LAYERS OF LTMS

The MoS2 occurs naturally as molybdenite. A crystal of molybdenite is illustrated in Fig. 2. Single crystals of MoS2 can also be grown synthetically by
vapor transport. However, synthetically grown MoS2 may occur in the 3R
polytype as opposed to molybdenite, which grows naturally in the 2H
polytype. The MoS2 can also be prepared by many methods described later.
Temperatures of preparation range from room temperature to temperatures
above 10008C. Low temperature preparation can lead to highly folded and
bent layers, termed the rag structure as shown in Fig. 3 (8). The ubiquitous
rag structure is most commonly found in MoS2 catalysts as described later.
The MoS2 particle shown in Fig. 3 is several layers thick and more than
100 nm across. However, catalytically stabilized MoS2 particles are often
single layers and approximately 5 to 10 nm across. Higher temperature
methods of preparation of MoS2 often produce Russian doll structures as
seen in Fig. 4 (9). These structures are often referred to as nested nanotubes
of MoS2 and have also been reported to occur in ReS2 (10). Nanotubes of MoS2
potentially have many useful catalytic applications as well as applications in
hydrogen storage and lubrication. However, a closer look at Fig. 4 will show
that the nanotubes are fractured and incomplete. It is not known at this
time if single walled nanotubes of MoS2 can be produced and much effort is
being expended to synthesize them since many important applications of
these materials can be envisioned. When considering single layer or single
walled nanotubes of MoS2 or other LTMS, we must consider the flexibility
inherent in the layers and to what degree the layers can be bent and still
maintain chemical integrity. The bending and compression of trigonal prisms
of Mo and S will limit the diameter to which the layers can be closed to form
nanotubes. One should note that carbon nanotubes are formed when a
change in stereochemistry is effected by introducing five membered rings

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Catalytic Properties of SLTMS

Figure 2: Naturally occurring crystal of Molybdenite (MoS2). The crystals are approximately
2 cm across.

into a six membered ring lattice. Such a change in symmetry has not yet been
reported in the LTMS family.
However, the LTMS do exhibit dynamic flexibility to a certain extent.
Although the catalytic literature involving TiS2 is very limited (11), the

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

R. R. Chianelli et al.

Figure 3: MoS2 in the rag structure (8).

material has been studied extensively for application as a battery cathode in Li


nonaqueous batteries. During discharge the TiS2 layers must open to accommodate entering Li cations as indicated in Fig. 5. This intercalation process
can also be achieved chemically and the process has been observed dynamically
in small single crystals using an optical microscope over a period of several
hours (12). In the dynamic intercalation process, the layers are bent and
stretched as shown schematically in Fig. 5. The moving front in the crystallite
is shown in Fig. 6. The process can also be followed using dynamic X-ray diffraction techniques (13). The X-ray study showed that the crystals would anneal to
their ordered state upon stopping the intercalation process. These studies
demonstrate that the LTMS have a degree of flexibility. In fact, in the Li intercalation studies for battery application, it was shown that the stress of bending
and straightening of the layers was a major factor in limiting the operation of
the cathodes at high rates. It is also concluded from these studies that the
limiting diameter of potential single-layer nanotubes of the layered LTMS
cannot be smaller than about 20 nm without stereochemical modification,
although more work is required to refine this point.

4. SYNTHESIS OF SINGLE LAYERED TRANSITION


METAL SULFIDES
The synthesis and properties of the LTMS and their intercalation compounds
have been recently updated (14). The MoS2 occurs naturally as described

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Catalytic Properties of SLTMS

Figure 4: MoS2 in the Russian doll structure (9).

previously. The WS2 and ReS2 do not occur naturally and must be synthesized
in the laboratory. Historically, the LTMS have been synthesized from the
elements at a temperature above approximately 4008C and single crystals
may be grown by various techniques, vapor transport being the most
common method (15). The high temperature growth methods produce crystalline LTMS that are excellent for fundamental studies of catalytic properties,
but these methods are of little use in practical application of LTMS catalysts.
Thus, low temperature methods have been traditionally used. The most
common method used to produce commercial catalysts is the aqueous impregnation of alumina supports using ammonium molybdate compounds. These
preparations are described in hundreds of patents and publications spanning
the long history of LTMS catalysis. An exhaustive review of these techniques

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

R. R. Chianelli et al.

Figure 5: Intercalation of Li cations into TiS2 from: Optical studies of transition metal
chalcogenides (12).

is found in Ref. 1. Further discussion of these techniques is beyond the scope of


this review, which concentrates on the synthesis of SLTMS and their catalytic
applications. However, more specific preparations of the SLTMS are described
in specific sections related to catalytic applications. A systematic method of producing transition metal sulfide (TMS) materials including SLTMS using low

Figure 6: Intercalating single crystal of TiS2 from: Optical studies of transition metal
chalcogenides (12).

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Catalytic Properties of SLTMS

temperature precipitation from nonaqueous solvents (16) was reported. This


produced a wide variety of the structures reported earlier including poorly
crystalline, amorphous, glasses and SLTMS. The SLTMS could also be
produced as dispersions in nonaqueous solvents such as dimethyl formamide.
This method and the properties associated with the materials produced have
been extensively reviewed (17).
In the exfoliation method, crystalline MoS2 is first reacted with
n-butyllithium to form LixMoS2, and then the LixMoS2 is reacted with water
in the presence of a surfactant producing single layers of MoS2 (18). The
single layers are produced as the reaction product of lithium with water, H2
escapes forcing the layers apart. The single layers can then be dispersed in
other solvents or flocculated, reassembling as new intercalation compounds.
For example, naphthalene can be inserted between the layers they reassemble
(19). In another variation of this method, single layered MoS2 is produced
directly in alkyl-amine melts using Mo(CO)6 and a sulfiding agent (20). This
method and related methods are the basis of new and interesting chemistry
that will produce novel and useful materials.

Figure 7: (a) Unoxidized crystal of MoS2 and (b) oxidized crystal from: The reactivity of MoS2
single crystal edge planes (23).

10

R. R. Chianelli et al.

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

5. CATALYTIC PROPERTIES OF LTMS AND SLTMS


EDGE PLANES
The anisotropic structure of the LTMS has been described previously. Though
the pervasive nature of the SLTMS in catalytic application has only been recognized recently, the importance of the anisotropy in catalytic reactions has long
been recognized. It was the original work of Tauster et al. using the technique
of oxygen chemisorption that pointed to the role of MoS2 edge area in catalytic
reactions (21). This led to the reemergence of the importance of the earlier work
by Thomas on the chemical nature of MoS2 edges (22). This elegant work
opened a series of optical and other studies on MoS2 single crystals in which
chemical reactions and their subsequent effect on the edges of an MoS2
single crystal could be observed directly. In Thomas original work the effect
of oxidation on the edges and defects in MoS2 is easily seen. This work was
repeated and optical micrographs of the eroded MoS2 crystallites can be
seen in Fig. 7 and schematically in Fig. 8. This work also added an extended
study of the MoS2 crystallites using Auger and scanning electron microscope
(SEM) techniques. These techniques clearly showed the chemical interaction
of oxygen, carbon, Co, and other elements with the MoS2 edges (23). The
Auger maps of Co on the edges of MoS2 crystallites are particularly striking
and important to the crucial problem of promotion by Co and Ni. An example
is shown in Fig. 9. Tanaka and Okuhara demonstrated the chemical reactivity

Figure 8: Schematic diagram of the oxidation of the MoS2 single crystal shown in Fig. 7.
Numbers on the crystal faces are in mm.

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Catalytic Properties of SLTMS

Figure 9: Auger elemental spatial mapping of Co on MoS2 single crystal edges from: The
reactivity of MoS2 single crystal edge planes (23).

of the edges in a dramatic way by studying the reactivity of single crystals cut
into pieces (24).
Classical high vacuum studies of MoS2 single crystals are virtually
nonexistent because the compound tends to lose sulfur under high vacuum
conditions and because Mo single crystals tend to form disordered and

11

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

12

R. R. Chianelli et al.

amorphous sulfides (25) when covered with sulfur. Additionally, it has been
shown that the basal planes of MoS2 single crystals are highly inert when
well-crystallized leading to the necessity of the studies described in the
previous paragraph (26). An alternate approach to studying the edges of
MoS2 single crystals was presented by Roxlo, et al. (27). In this approach,
single crystals of MoS2 were etched by a lithographic technique to produce
pyramids of MoS2. The pyramids where then studied with photodeflection
spectroscopy (PDS) (28). These studies associated optical absorption above
the band edge of MoS2 with dangling bonds that give rise to the catalytic
activity. A good correlation to this absorption in several catalysts was
obtained. This and related work gave insight into the actual chemical nature
of the defects that give rise to the catalytic activity of MoS2 edge sites as
described more thoroughly in the next section.
In addition to the oxygen chemisorption and PDS correlation to HDS
activity as previously described, other correlations to the activity of the edge
planes have been reported. For example, both NMR and ESR give good correlations to HDS activity (29, 30). The ESR measurements on a variety of MoS2
samples gave a straight-line correlation for the HDS activity in model
compound studies. Approximately 10  1018 spins/gram of catalyst were
reported in good agreement with the calculated edge area of the samples.
Magnetic susceptibility and ESR were measured together at 10 GHz on molybdenum sulfide catalytic materials. The densities of paramagnetic centers
inferred from the two types of measurements differ by an order of magnitude
(approximately 1 and 0.1 mol%, respectively). This demonstrated that the
ESR measurements probe a subset of the magnetic species detected by the susceptibility measurements. This result led to a model that suggested pairing of
Mo atoms at the edge and single Mo atoms giving unpaired spins at the corners
of the particles. Thus, the catalyst particles are electronically conducting at the
edges. This idea has been recently confirmed in STM studies as described later.
The property of electron quantum confinement due to the presence of small
sheets of MoS2 is important to photocatalytic effects described later, but also to
the presence of small sheets of MoS2 in supported HDS catalysts. This effect
was described by Garoff et al. (31) through application of Mie scattering
theory. The measurement of the optical response of materials occurring as
finely divided powders is difficult due to the complex interaction of the optical
wave with the material. Mie theory was used to study the optical properties
of composites such as supported MoS2 catalysts. Experimental and theoretical
techniques were developed for characterizing the optical properties of these
materials. The results need to be exploited further because the techniques
give direct information regarding the electronic properties of the SLTMS.
It should be noted, however, that all the work described earlier was performed on stacked MoS2 layers and though catalytically relevant still leave
open questions regarding the precise nature of the SLTMS occurring in

Catalytic Properties of SLTMS

single layers. This means that in this article several fascinating and important
catalytic applications and studies are not addressed. For example, the synthesis of alcohols by alkali promoted MoS2 catalysts probably requires
multiple layers in a stacked sequence (32).

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

6. PHYSICAL AND CHEMICAL STRUCTURE OF LTMS


AND SLTMS EDGE PLANES
When discussing what is known about the structure of single layers it is essential to distinguish between the catalyst precursor and the stabilized
catalyst. The discussion that follows is limited to MoS2 catalysts that are
used in hydrogenation processes containing sulfur. These applications are currently the major use of these catalytic materials. The scheme described later
will change according to the catalytic environment. The simple scheme that
follows indicates that the edge plane chemistry changes after sulfidation
and after stabilization in the catalytic environment.
Sulfidation ) MoS2x 0 , x , 0:5
catalytic conditions ) MoS2x Cy 0 , y , 1
The values of x and y grow as the particle size decreases. A large amount of
information exists for the sulfided precursors and much less for the stabilized
catalyst state. The fact that the MoS2 edge planes are the catalytically important sites is well established. What are the chemical states of these edge plane
sites? The answer depends on the conditions under which the observations are
made. For example, Chang and Chan (33) observed that after sulfiding MoS2
catalyst precursors disulfide bonds were observed using infrared (IR) and
Raman techniques. They observed that catalytically active, poorly crystalline
MoS2 exhibit sharp infrared bands at 385 and 470 cm21. However, broad
bands at 287, 335, 373, and 522 cm21 appear after H2S treatment indicating
the presence of polysulfide bands. The intensity of the IR peaks increased as
the MoS2 edge area increased (33). This would lead to a composition that can
be described as MoS2-x(S S)0.5y . A simulation of this state is shown in
Fig. 10. In this example, the hypothetical cluster would have the stoichiometry
MoS1.88(S S)1.12. The polysulfide bonds could be placed in other locations,
illustrating the complexity of defining the MoS2 catalytic edge planes.
In Fig. 11 we see another hypothetical MoS2 cluster after the edge planes
are stabilized in a catalytic hydrodesulfurization environment. In this case,
the catalyst is sulfur deficient and contains carbon in a carbide environment.
The stoichiometry in this case is MoS1.88C1.12. Although, the exact structure of
the edge carbide is not known the carbiding of the MoS2 edge plane is supported
by synchrotron studies of catalysts stabilized in HDS conditions (34). In this
study, carbon edge X-ray absorption studies clearly indicated through the

13

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

14

R. R. Chianelli et al.

Figure 10: Cerius2 simulation of MoS2x. A small cluster with the stoichiometry MoS1.88 (SS)1.12.

Figure 11: Cerius2 simulation of MoS2-xCx.

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Catalytic Properties of SLTMS

measurement of Auger electrons emanating from the catalyst surface as a


function of X-ray energy, that the surfaces of MoC and the stabilized MoS2
catalysts were identical. This is an example of the surface of a working
catalyst and indicates that understanding of catalyst structure/function
relations requires the study of such stabilized catalysts and that the study of
precursors, though useful in understanding the chemistry of the catalytic
material, is not sufficient.
A strikingly beautiful image of a single layer MoS2 catalyst precursor
created under vacuum and then sulfided can be found in the work of H.
Topse (35). The atomic force microscopy (AFM) image of a single triangular
layer of MoS2 may be seen in Fig. 12. The single layer crystallite is approximately 3 nm wide and was prepared on a Au(111) substrate by depositing the
metal in an atmosphere of H2S (1  10 6 mbar at 400 K) and then annealing
(673 K for 15 minutes). The triangular shape was unexpected but, as described
later appears to be the stable state for single layer MoS2 hydroprocessing
catalysts. Sulfur vacancies were also imaged after treatment with hydrogen
by treating the catalyst with atomic hydrogen. The edge chemistry of these
materials is complex and flexible as expected for a catalyst, but the basic
morphology of the MoS2 illustrated in Fig. 12 best describes a model for
subsequent fundamental studies of the catalytic chemistry of the SLTMS.
However, one should consider that operating SLTMS are generally highly
disordered, thus introducing another level of complexity. Detailed theoretical
studies of the physical and chemical structure of the SLTMS are described in
the next section.

7. THEORETICAL STUDIES OF THE PHYSICAL AND


CHEMICAL STRUCTURE OF SLTMS EDGE PLANES
It has been well established that the catalytically important electronic states at
the edge of the MoS2 layers are dangling d electron states. This was established by electron spin resonance (ESR) and magnetic susceptibility studies
(36). These ESR studies measured 10 20  1018 spins per gram of MoS2
catalyst and this correlated well with the oxygen chemisorption studies
described earlier. A paramagnetic model was assumed that described the
edges of the MoS2 as electronic, paramagnetic conductors. The dangling
bonds of catalysts in the precursor state (sulfided) were studied using XPS
and UPS photoemission techniques (37). The electronic energies measured in
this study were:
dz2  2:2 eV nonbonding d electrons
dp  39 eV bonding Mo 4d and S 3p
s  13:5 eV non bonding S 3s

15

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

16

R. R. Chianelli et al.

The study further concluded that the tail above the dz band was due to dangling
bond catalytic defects that reversibly chemisorbed/physisorbed oxygen and
that these edge defects were possibly dimers that were electronically conductive. Recent theoretical studies confirm and expand this idea (38).
In the past 20 years theoretical computing has become easier and theoretical studies of catalysts have become more and more common. This is particularly true in the case of the subject materials of this article. Density
functional theory (DFT) has become almost routine as applied to MoS2 catalytic
particles. At this writing the theoretical work has concentrated primarily on
the structure of MoS2 precursors in the sulfided or activated state. Many
calculations on MoS2 layers have been done. The most useful are those that
are accompanied with complementary surface and analytical techniques. A
recent example of this approach is the result of the efforts of the research
group of and colleagues Klier (39, 40). Stable edge periodic structures were
predicted with edge stability found to be in the order (1 0 1x) (most
stable) . (1 . 2 1x) . (1 . 2 1 0) . (1 0 . 1 0), where the inclination index x 3
or 4 (inclination index angle of cutting the plane). A large relaxation
energy associated with the reconstruction of the edge from ideal geometry of
a cut through the 2H-MoS2 crystal: for the (1 0 . 1 x) edge, energy of 0.63 eV
per MoS2 molecular formula is released upon a concerted movement of
exposed S atoms that increases the coordination of the edge Mo atoms from 4
to 5. The calculated states were also metallic in agreement with previous
studies. However, calculations on single layers of MoS2 were not included
and the results reported are for stacked crystallites. These studies also
included detailed calculations using DFT methods of MoS2 monomers,
clusters, and edges in periodic structures as they activate hydrogen.
Hydrogen was found to bind in stable configurations to Mo or S atoms or in
bridging positions across two Mo atoms. From these calculations the following
vibrational frequencies were calculated:
Mo-H-Mo 1223 cm1
Mo-H 1860 cm1
Mo-S 2500 cm1
An example of the power of combining surface studies (STM) with theoretical
calculations (DFT) can be found in recent work by the Topse group (41).
Single layers of MoS2 were synthesized and studied using STM techniques.
The STM studies were performed in an environment of hydrogen and thiophene with the surprising result that thiophene adsorbed on full sulfided
edges. The binding sites are one dimensional metallic brim sites (metallic
edge site) as described by DFT calculations. This result is in contrast to the
common wisdom that coordinately unsaturated sites (CUS) sites are required

Catalytic Properties of SLTMS

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

for the molecular binding to occur. Reconstruction of the edge sites is also
shown in these studies along with remarkable details regarding the formation
of vacancies, the binding of hydrogen and thiophene, and the mechanism of
desulfurization under STM conditions. The combination of DFT techniques
and surface techniques is the way to future detailed information regarding catalytic pathways. For this information to be related to commercial catalysts, the
catalyst must be studied after it has been stabilized under operating conditions
and the structural role of carbon explained.

8.

SINGLE LAYERS IN COMMERCIAL CATALYSTS

As described previously the reactivity of MoS2-based catalysts depends on


disordered edge planes presenting themselves to the reactants in a catalytic
environment. The edge planes are highly active, unlike the basal planes that
are thought to be inert. The total activity depends on the number of these
sites present in a stabilized catalytic environment. Basic studies prove that
the selectivity of unsupported MoS2 catalysts are affected by the stacking
height of the MoS2 slabs in the Rim and Edge model proposed by Chianelli
et al. (42). In such a model, the rim sites are active for hydrogenation reactions
and C S bond rupture while edge sites are active only in C S rupture (see
Fig. 13). A freshly prepared active catalyst presents stacks of four to five
slabs. Much of the HDS literature has attempted to rectify the rim/edge
theory with selectivity studies done on commercial alumina-supported catalysts. However, the original rim/edge theory did not include the promotion
effect on selectivity and, furthermore, most studies have been done on commercial catalysts that are not stabilized in a catalytic environment. Recent studies
show that when a commercial catalyst is stabilized in an operating catalytic
environment, only single layers exist as described later. This fact has been
missed until recently and is a crucial concept for understanding and optimizing
these catalysts. In addition, it has become necessary to reevaluate and
understand the selectivity of promoted single layer commercial catalysts at a
fundamental level.
Numerous techniques have been extensively used in an attempt to understand the structure of the catalyst active phase including high resolution transmission electron microscopy (HRTEM) and X-ray absorption spectroscopy
(XAS). However, the highly anisotropic layered stacked structure has limited
the information that can be drawn from the catalytically stable state of an
HDS catalyst. Furthermore, the interaction of the support with the active
phase adds to the difficulty in determining catalyst dispersion. Perhaps one
of the leading techniques used to average slab lengths and mean values for
the stacking numbers is HRTEM. However, its ability to detect the MoS2
slabs strongly depends on how the slabs are bonded to the support (43).
Edge-bonded layers are clearly visible whereas thin flat slabs of MoS2 are

17

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

18

R. R. Chianelli et al.

Figure 12: AFM micrograph of single layer of MoS2 on gold substrate (35).

Figure 13: Rim and Edge Theory (42).

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Catalytic Properties of SLTMS

simply missed. The XAS measurements underestimate the dimensions along


the basal plane because ordered sulfide particles are assumed in the fitting
procedure while they are disordered and distorted in real catalysts (44).
Structural analysis of X-ray diffraction patterns on the other hand, probes
all orientations of the MoS2 slabs and gives an unbiased picture of the
catalyst morphology.
In a recent study by Perez De la Rosa et al. (45) three commercial Co/
MoS2/Al2O3 catalysts were analyzed using synchrotron radiation X-ray diffraction. The catalysts differ by the length of time of use: the first run for one week,
the second run for one month in a pilot plant, and the third run in an industrial
plant for four years. The widths of the diffraction peaks were used to estimate
the size of the lattice coherence normal to the diffraction vector. Liang and
coworkers (46) established a procedure for quantitatively determining the
size of the MoS2 slabs from the analysis of the diffraction pattern of poorly
crystalline MoS2 (illustrated in Fig. 14). Following their procedure the (002)
peak was used to determine the height and the (110) peak to determine the
diameter of the MoS2 slab. The line broadening analysis was complemented
by a full scattering model. The full scattering model evaluates the area of the

Figure 14: Typical X-ray diffraction profile of the poorly crystalline MoS2 phase. The positions of
the main diffraction peaks are indicated by arrows (45).

19

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

20

R. R. Chianelli et al.

(002) peak and the diffuse scattering area under the (002) peak to determine
the fractions of stacked and unstacked layers. The relative proportions of
stacked and unstacked layers were then directly obtained from the X-ray
diffraction (XRD) data. Their results were surprising and unexpected. After
four years of industrial hydrotreating operations, the synchrotron scattering
patterns of these catalysts show that they are completely destacked (see
Fig. 15). The catalyst consists of single layers of MoS2 with a minor loss of
activity. As the pressure increased from 5.5 MPa to 7.8 MPa a strong
decrease in stacking was observed. Therefore, high pressure applied during
the hydrotreating conditions was determined to be the main cause for the
destacking effect observed in the spent commercial catalysts under industrial
conditions. This result was also seen in freshly prepared catalyst run in dibenzothiophene (DBT) reported in the same paper (46).
A materials study with similar pressure-crystallization effects was
reported by Peng et al. (47) where the slab stacking decreased with increasing
pressure inside the reactor during the hydrothermal synthesis of MoS2 in an
autoclave. The authors suggested that perhaps other experimental parameters
of the HDS process, such as the hydrocarbon pressure or concentration could
contribute to the destacking effect. The formation of multilayered stacks
through van der Waals forces seems counterbalanced by the strong interaction

Figure 15: The low-angle X-ray synchrotron patterns for a commercial CoMo/Al2O3
catalyst at the three different stages of catalytic life: one week (A), one month (B), and
four years (C) (45).

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Catalytic Properties of SLTMS

Figure 16: Destacking effect under the hydrodesulfurization process: A) sulfided precursor,
B) hydrogen activated catalyst, C) stabilized single layer catalyst.

of adsorbed hydrocarbon molecules that provide stability to single MoS2 layers


(see Fig. 16). If a strong interacting support is present the destacking effect is
accentuated. This result also confirmed a recent study by Glasson et al. (48)
that showed that less stacked CoMo/Al2O3 catalysts could be obtained if they
were pretreated with gas oil or other oils. Their treatment at atmospheric
pressure found that hydrocarbons are able to enhance the destacking
process. However, applying high pressure will favor the stabilization effect
by increasing the interaction between hydrocarbons and MoS2 single layers.
In retrospect, this point seems obvious in that if hydrogen intercalates at
high pressure, then the layers would simply slide apart with time. This fact
however, emphasizes the importance of studying stabilized catalyst to
develop fundamental insight. Further support for the single layers being the

Figure 17: Optical micrograph of Mo microcat catalyst dispersed in Cold Lake crude after
hydroconversion reaction.

21

22

R. R. Chianelli et al.

stable catalytic entity in HDS catalysis is found in a recent paper that describes
the morphology of carbon-supported WS2 catalysts (38). The paper also reports
images of triangular nanoclusters.

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

9. SINGLE LAYERED TRANSITION METAL SULFIDE (SLTMS)


CATALYSIS IN HEAVY OIL PROCESSING
In the previous section we described catalysts that commercially available in
the form of solid pellets and usually are used in various fixed bed reactor
systems. However, in this section we describe certain finely divided, transition
metal based catalysts that are used as slurries in heavy crude oils (e.g.,
Canadian bitumen) or the residual fraction thereof (i.e., material typically
boiling above 5668C) to catalyze the hydroconversion upgrading of these
heavy materials. Hydroconversion upgrading denotes conversion of oil
boiling above 5668C to liquid products boiling below 5668C, along with significant removal of contaminants such as sulfur and metals. A major role played by
catalysts in all slurry technologies is suppression of coke formation in the
hydroconversion step that operates at thermal cracking temperatures
(i.e., .4008C). Catalyst or catalyst precursor can be added to the feed by one
of several methods that include addition as a finely divided powder, disperion
in feed as an emulsion of a water solubilized metal compound, or by addition
as an oil soluble metal compound.
This general area of technology dates to the 1920s when Pier added catalyst
slurry to the Bergius high-pressure process for hydroliquefaction of coal (49).
The apparent first application of transition metal slurry catalysis for petroleum
was reported by Zorn in 1932 (50). However, it was not until about 1960 that
there was a resurgence of interest in slurry catalysts for treating heavy
petroleum feeds. Numerous patents and publications have appeared since
1960 and research continues to the present day. Ten slurry-catalyst hydroconversion processes have evolved from this body of research (Table 1).
Patents that cover these various processes generally provide considerable
latitude with respect to the catalyst metal or metals that are used, as well as
catalyst (or additive) concentration on feed. Selections quoted in Table 1 are
best estimates based on references cited.
There is little information in the literature concerning the SLTMS
character of the catalysts that are used in the various processes cited in
Table 1. The two exceptions are for the catalysts that are used in the
Microcat-RC process and in the ENI-EST process. Both of these processes
operate with catalysts that contain single layer MoS2 platelets. A possible
explanation for SLTMS formation in MoS2 slurry systems, i.e., coordination
with asphaltenes and/or resins in heavy feeds, is discussed in the following
section on Microcat technology.

Catalytic Properties of SLTMS


Table 1: Slurry catalyst hydroconversion processes
Process

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Veba Combicrackinga
Aurabonb
Canmetc
HDHd
Microcat-RCe
SOCf
(HC)3g
ESTh
CASHi
Tervahl-Cj

Catalyst/Precursor

Conc. on feedk

Reference

Coal char or iron ores, etc.

1 3 wt%

[51]

VSx/VOSO4 (e.g.)
Fe7S8/FeSO4 . H2O (coal
optional)
Natural ore, e.g., iron laterite
MoS2/H3PMo10O34
MoS2 (Carbon black)
MoS2/Mo 2-ethyl hexanoate
MoS2/Mo naphthenate
MoS2/ammonium oxy sulfides
and heptamolybdate
MoS2/various salts of Mo,
including phosphomolybdic
acid and molybdenum blue

2 6 wt%
1 2 wt%

[52]
[53, 54]

1 3 wt%
100 300 wppm Mo
,1.0 wt%
,150 wppm MoS2
Not availablel
Not availablem

[55]
[56, 57]
[58]
[59]
[60, 61]
[62]

,100 wppm Mo

[63, 64]

Veba OEL process.


Process developed by Universal Oil Products.
PetroCanada process developed at the Energy Research Laboratories of the Canada Centre
for Minerals and Energy Technology.
d
Inteveps process, HDH is the acronym for Hydrocracking, Distillation, Hydrotreating.
e
ExxonMobil process for residuum conversion where RC denotes resid conversion.
f
Super Oil Cracking process developed by Asahi Chemical Industries, Japan Energy Company
and Chiyoda Corporation.
g
Process developed by Alberta Energy (formerly AOSTRA). The acronym stands for
HighConversion HydroCracking HomogeneousCatalyst.
h
Eni Slurry Technology is a process developed by EniTecnologie.
i
CASH is an acronym for Chevron Activated Slurry Hydroprocessing.
j
A catalytic hydrovisbeaking process shared by Total, Elf Acquitaine, and Institut Francais du
Petrole.
k
Catalyst concentration in reactor liquid can be substantially higher than that of feed as a
result of volatilization of hydrocarbons and/or recycle of catalyst.
l
Mo concentration in reactor liquid, maintained via recycle, greater than 1000 wppm. Makeup
Mo concentration in feed not specified.
m
Mo concentration in reactor liquid, maintained via recycle, is in range of 0.1 to 1.0 wt%.
Makeup Mo concentration in feed not specified.
b
c

a. Microcat Technology
The preferred catalyst entity for ExxonMobils slurry-catalyst hydroconversion process (MRC process) (65) for upgrading heavy crudes and residua
is a nominal one-to-two micron diameter particle (Fig. 17) that is comprised
of single layer molybdenum sulfide clusters associated with a hydrocarbonaceous matrix (Fig. 18) (65 67). Such catalyst properties are important for
this hydroconversion process wherein liquid phase free-radical hydrogenation
is used to help eliminate formation of coke (56, 65). Small particle size provides
excellent dispersion of catalyst in reactor liquid and single layer MoS2
distribution appears to provide optimum utilization of molybdenum (65). The
net effect is that only trace amounts of Mo, 100 to 300 wppm based on

23

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

24

R. R. Chianelli et al.

Figure 18: Transmission electron micrograph of small particle Microcat catalyst prepared
from PMA (phosphomolybdic acid) dispersed in Cold Lake crude (66).

hydroconversion feed, are required to run the process, and catalyst can be used
) in the transmission
on a once-through basis. Short dark lines (50 to 100 A
electron micrograph (Fig. 18) (66) represent edge views of single layer MoS2
,
clusters. Some regions show layering but the interlayer spacing is above 10 A
reported for pillared, crystalline MoS2 (68). The dark gray amornot the 6.17 A
phous area of the micrograph is thought to be the catalysts hydrocarbonaceous
matrix.
The MRC process operates in the 420 4508C range under a total pressure
of 10 15 MPa. Conversion is driven by thermal cracking as it is in coking
technologies, but coke formation is essentially eliminated. Control over coke
formation is attributable not only to the catalyst but also to reactor design
and operating conditions. Conceptually, the catalyst acts as a hydrogen
transfer agent to cap intermediate free radical moieties in the liquid phase,
thereby inhibiting secondary radical reactions leading to formation of coke
precursors and gas (Fig. 19) (65 67). Note the large C5-5668C liquid yield
advantage for MRC over delayed coking.

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Catalytic Properties of SLTMS

Figure 19: Catalytically controlled thermal cracking.

The Mo Microcat catalyst described here is one of a family of Microcat


catalysts that have been prepared from Groups IB, IVB, VB, VIB, VIIB, and
VIII metals (57). These catalytic materials can be formed in situ in the
process feed, typically a heavy crude oil or the 5668C fraction of the crude
oil, by heating a mixture of feed and suitable catalyst precursor preferably in
the presence of sulfiding agent and hydrogen. Preferably, a concentrate that
contains the catalyst metal is prepared in a portion of process feed, and the concentrate is metered into the main feed stream to provide the desired level of catalytic metal in reactor liquid. For the molybdenum-based system, the Mo
component of the concentrate can be the catalyst precursor, a presulfided precursor, or a preformed catalyst (69, 70). One of several molybdenum compounds
can be used as catalyst precursor. However, a phosphomolybdic acid (PMA) is
preferred. Phosphomolybdic acids are readily prepared via reflux of puregrade MoO3 with dilute phosphoric acid at the desired P/Mo atom ratio (71).
Moreover, when an aqueous solution of PMA is blended with heavy crude oil,
the acid reacts with organic bases in the oil, thereby leading to an essentially
homogeneous dispersion of PMA. The aqueous PMA used in most of this
work had an empirical formula of H3P1.25 Mo10O34.
Studies on catalyst preformation, i.e., preparing catalyst concentrate in a
reactor separate from the hydroconversion reactor, led to an improved understanding of the chemistry of catalyst formation. For example, in a preparation
sequence using PMA dispersed in Cold Lake crude, phosphomolybdate was
first converted to MoS3 (likely present as thiomolybdate) by treating with
hydrogen sulfide at 150 2008C. The crude oil plus MoS3 was homogeneous
at this point, but upon heating to 320 3308C (with or without hydrogen), a
dispersion of oil-insoluble catalyst appeared (Fig. 20). The temperature at

25

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

26

R. R. Chianelli et al.

Figure 20: Threshold temperature for Mo Microcat formation from PMA (phosphomolybdic
acid) dispersed in Cold Lake crude.

which catalyst formation occurs is consistent with the range reported for the
decomposition to (NH4)2MoS4 and then to MoS12.
The hydrocarbonaceous matrix that forms is not a pyrolytic coke. The
temperature of formation is too low for reactions leading to coke. Furthermore,
when Cold Lake crude oil is subjected to the same treatment conditions without
to molybdenum present, solids are not obtained. Thus, it appears that single
layer MoS2 clusters react or coordinate with components of the feed to form
the hydrocarbonaceous matrix. An elemental assay of Mo Microcat catalyst
formed in a Cold Lake crude shows that the hydrocarbonaceous matrix has a
composition similar to that of the heptane insoluble asphaltene fraction of
the crude (Table 2). The hydrogen to carbon atomic ratio is as expected, as is
the high content of heteroatom impurities. The nature of association between
single layer MoS2 clusters and hydrocarbonaceous matrix is not known, but
is likely to involve coordination with organo-sulfur sites within the matrix.
Catalyst formed from molybdenum naphthenate dissolved in Cold Lake
crude oil yields a catalyst of comparable activity to that obtained with PMA
(57). In the initial step of catalyst formation, molybdenum naphthenate decomposes (starting at about 1508C) to yield a dispersion of MoO3 in oil. From that
point on, catalyst formation steps are thought to parallel those of the PMA
system. Catalyst infrastructure is similar to that of the PMA-derived

Catalytic Properties of SLTMS


Table 2: Mo Microcat catalyst assay: Nature of hydrocarbonaceous
matrix, catalyst formed with PMA in Cold Lake crude (67).

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Component,
wt%
Mo
C
H
S
N
Ni V
H/C atom ratio

Catalyst

Matrix
(normalized)

C7
asphaltene

10.9a
65.5
5.4
16.8
1.2
,0.1
1.0

0.0
80.2
6.6
11.8
1.5
,0.1
1.0

0.0
81.9
7.9
7.5
1.2
0.1
1.15

.99% MoS2 by ESCA.

) intercatalyst but with more ordering of layers, albeit with very large (20 A
layer spacing (Fig. 21) (66). Without further characterization, it is difficult to
say whether heptane-insoluble asphaltenes are the sole contributor to formation of the catalysts hydrocarbonaceous matrix. Resins (polar aromatics)
are also possible candidates. In view of this uncertainty, Conradson carbon
(CCR) was chosen as a feed parameter for study of feed composition effect on

Figure 21: TEM of Mo Microcat catalyst formed from Mo naphthenate dissolved in Cold Lake
crude oil (66).

27

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

28

R. R. Chianelli et al.

catalyst properties. The CCR is a measure of molecules in heavy crude or


residuum that will form coke upon thermal cracking (72). Asphaltenes and
resins are the principal contributors to CCR (73).
Several catalysts were prepared in a hydrocarbon medium in which various
blends of Cold Lake crude (11 wt% CCR) and a Cold Lake derived vacuum gas
oil (1.0 wt% CCR) were used to vary the amount of CCR present for catalyst
formation. Molybdenum naphthenate was used as the catalyst precursor.
Catalysts were formed by heating at 3858C for 30 min with 5 molar% H2S in
hydrogen under 12 MPa total pressure (74). Results of these experiments
(Fig. 22) show that catalyst particle size and surface area vary inversely with
the CCR/Mo atomic ratio, where CCR was given the atomic weight of
carbon. Catalysts that are preferred for controlled thermal cracking, the
small particle size single layer MoS2-containing catalysts, are formed at high
ratios (e.g., 3000:1 or higher). High surface area, larger particle catalysts
with layered MoS2 are favored by low CCR/Mo ratios. One conclusion that
can be drawn from this correlation is that the hydrocarbonaceous matrix is
needed in a high concentration to stabilize single layer MoS2. This conclusion
is further illustrated by assays of catalysts formed at high and low CCR/Mo
ratios (Table 3). The TEM of high surface area Mo Microcat (Fig. 23) indicates
a fairly amorphous composition. There are islands of highly folded MoS2 as
described earlier, but with no more than three to six layers.
High surface area Mo catalysts, because of larger particle size, perform
poorly for coke suppression when used at low Mo concentrations on feed

Figure 22: Effect of Conradson carbon to Mo ratio on catalyst particle size and surface area.

Catalytic Properties of SLTMS


Table 3: Molybdenum Microcat catalysts (prepared by oil in-situ
decomposition of Mo naphthenate).

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

CCR/Mo atom ratio


Particle size range, microns
Surface area, m2/g
Elemental assay, wt%
Carbon
Hydrogen
Molybdenum
Sulfur

3000:1
1 2
20
72.2
6.0
5.8

11:1
50 100
335
11.8
1.1
47.1
35.0

(i.e., 100 300 wppm Mo). On the other hand, when used at concentrations in
the 1 3 wt% range and above, the high surface area catalysts provide coke
suppression as well as enhanced rates of 5668C conversion, sulfur removal,
and Conradson carbon conversion (67). Because of cost, the use of effective
amounts of high surface area Mo catalysts for heavy crude hydroconversion
requires provisions for catalyst recovery and recycling. High surface area
Microcat catalysts have been prepared with metals from Groups IB, IVB,
VB, VIB, and VIIB (76).

Figure 23: Electron micrograph of high surface area Mo Microcat Catalyst prepared with
CCR/Mo atomic ratio of 11:1 (75).

29

30

R. R. Chianelli et al.

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

b. ENI-EST Catalyst
Another example of MoS2 single layer catalysis is found in work relating to
ENI-EST hydroconversion technology (61). This Mo-based catalyst was formed
in situ in sulfur-rich Belayin vacuum residue from molybdenum naphthenate.
An X-ray analysis of the catalyst recovered after a standard hydroconversion
test suggested the presence of monolayers of MoS2 or of crystallites with a
maximum of two or three layers. A TEM-EDS analysis indicated the
presence of isolated slabs of MoS2 and, to a lesser extent, slabs grouped in
. Several catalysts other
sheaves with parallel and regular spacing of 25 30 A
than the molybdenum-based catalyst were also screened. These catalysts
were formed in situ in the feed from oil soluble salts, including those of iron,
nickel, cobalt, vanadium, and ruthenium. Molybdenum was preferred based
on overall performance.
Based on findings for the Microcat and EST catalysts, it appears likely that
MoS2 catalysts used in the SOC, (HC)3 CASH, and Tervahl-C technologies
(Table 1) also contain single layer MoS2.

c. Related Studies
The direct liquefaction of coal using MoS2-based catalysts (77, 78) is also an
area where SLTMS catalysis may be involved, as well as in the hydroconversion of coal and heavy oil mixtures (79). Also, a study has been reported
where exfoliation/restacking of MoS2 crystallites was applied in exploratory
studies as a methodology for development of layered catalysts, especially for
coal liquefaction and hydrorefining of petroleum residues (80). Several department of energy (DOE) contract studies have been reported for MoS2 slurry catalysts, iron catalysts, and Mo promoted iron catalysts for the direct liquefaction
of coal as well as the hydroconversion of coal mixed with oil (81).
Overall, it appears that single layered Group VIB metal sulfides will
remain an important tool in catalytic hydroprocessing, particularly as more
of the worlds enormous reserves of heavy crude oils are needed to meet
energy demands.

10. PHOTOCATALYSIS BY SLTMS (PHOTOCATALYSIS


USING MOS2 NANOCLUSTERS)
Efficient photocatalytic destruction of toxic waste or splitting of water to
produce hydrogen requires a photocatalyst capable of being activated by
visible light. In the bulk form, MoS2 has a band gap corresponding to near
infrared wavelengths. However, because of quantum confinement on the
properties of nanosized MoS2 (n-MoS2) the optical gap is shifted into
the visible wavelength regime. Accompanying optical gap widening is the

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Catalytic Properties of SLTMS

favorable alignment of the valence and conduction bands with redox


potentials of interest. Thus, nanosized MoS2 photocatalyst can be driven
with visible radiation. Several studies have reported the photocatalytic
activity of bulk TMS and SLTMS (82) but few have studied these properties
on the nanoscale.
Initial studies in photooxidation of toxic wastes such as phenol using
n-MoS2 were performed by Thurston and Wilcoxon (83). They used different
sizes of n-MoS2 clusters (4.5 nm and 8 nm) to examine the effectiveness of
phenol destruction compared to the standard Degussa P25 TiO2 photocatalyst
(83). Using high pressure liquid chromatography (HPLC) they could monitor
the concentration of phenol as a function of time after illumination of the
n-MoS2 photocatalyst/phenol solution at 455 nm. Figure 24 shows a 225 nm
absorption chromatogram using 4.5 nm MoS2 where the phenol elutes at
14.6 minutes. Over an 8-hour period, the area of the phenol peak decreases
by 25% accompanied by the appearance of two phenol photooxidation
products: catechol (13.4 min) and a possible isomer of catechol (12.2 min). The
Degussa P25 TiO2 exhibited a similar reaction pathway for phenol photooxidation upon illumination with 365 nm radiation (Fig. 25). However, there was no
reaction upon illumination with visible light (Fig. 26). The active sites on the
MoS2 clusters corresponding to the active Ti (IV) sites of TiO2 are presumably

Figure 24: HPLC absorption chromatogram showing phenol destruction with visible light
(455 nm) using 4.5 nm MoS2 clusters as the photocatalyst (83).

31

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

32

R. R. Chianelli et al.

Figure 25: HPLC chromatogram showing phenol destruction with 365 nm UV illumination
catalyzed by Degussa P25 TiO2 (83).

the empty d-orbitals accessible at the Mo metal edge sites. Phenol absorption
and hole transfer could occur at these locations.
Other important requirements in photocatalysis (or in any catalytic
process) are photocatalyst regeneration and degradation resistance.
Figure 27 shows the absorbance spectrum (collected as part of an HPLC chromatogram) of 4.5 nm MoS2 clusters before and after photocatalysis (84). There
was no decrease in the area under the curves (representative of MoS2 concentration) with either visible or ultraviolet (UV) illumination. There are no shifts
in the characteristic excitonic peak positions for MoS2 before and after the
photocatalytic oxidation of phenol. This shows the true photocatalytic nature
of MoS2 nanoclusters, i.e., it is regenerated and does not exhibit any signs of
photodegradation.
Photooxidation of phenol was also possible using 8 10 nm MoS2 clusters
but not to the extent as of the 4.5 nm clusters. The extent of phenol photodestruction is represented in Fig. 28, which shows a plot of phenol concentration as
a function of time under visible illumination for 4.5 nm MoS2 clusters, 8 10 nm
MoS2 clusters, and Degussa P25 TiO2. There was a steady decline in the phenol
concentration with the 4.5 nm MoS2 (Fig. 24), some reduction with 8 10 nm
MoS2 (Fig. 25), and no change with the P25 TiO2 (Fig. 26). For the 4.5 nm

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Catalytic Properties of SLTMS

Figure 26: HPLC chromatogram of attempted phenol destruction using Degussa P25 TiO2
illuminated by visible light (455 nm). No decrease in phenol concentration is observed and no
photooxidation products develop (83).

MoS2 the phenol destruction rate could be replicated upon adding more phenol
after most of the phenol had been destroyed (indicated in Fig. 26). This is another
demonstration of the regenerative photocatalytic nature of these nanoclusters.
Quantum confinement effects were also observed during photooxidation
studies of pentachlorophenol (PCP) using even smaller (3 nm) sized MoS2
clusters (85). The PCP is a chlorinated aromatic molecule from the chlorinated
phenol family. This very toxic chemical is thought to originate from water treatment using chlorine in the presence of organic materials, sewage treatment
plants, and incinerators. Its slow natural degradation rate allows it to
persist in the environment. Direct photolysis of PCP leads to highly toxic byproducts such as octachlorodibenzo-p-dioxin (86). Many studies have shown the
total mineralization of PCP to CO2 and HCl using TiO2 (87, 88). However,
due to the large optical gap of TiO2 (3.2 eV), it is limited to excitation by UV
light, which is only 3% of the solar spectrum. Excitation with visible light
would allow for use of a larger part of the solar spectrum. Since MoS2
clusters are highly active under visible illumination, they are ideally suited
for visible light photooxidation studies of PCP.
Bulk CdS powder was used as a comparative standard to the MoS2 clusters
studied (68). Using HPLC the mineralization rate of the PCP could be

33

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

34

R. R. Chianelli et al.

Figure 27: Optical absorption of MoS2 before and after photocatalysis showing no
change (84).

determined. The decrease in PCP concentration as a function of time for two


sizes of MoS2 clusters (4.5 nm and 3 nm), bulk CdS powder, bulk MoS2
powder, and TiO2 illuminated with visible light is shown in Fig. 29 (85). Bulk
MoS2 and bulk TiO2 had no effect on the PCP under visible light radiation,
as expected. The bulk CdS powder standard did show some ability to slightly
mineralize the PCP. However, both sizes of MoS2 clusters exhibited substantially more activity than the CdS standard.
A dramatic increase in PCP destruction rate occurred for the 3 nm MoS2
compared to the 4.5 nm MoS2 (Fig. 29). Complete photooxidation of PCP
occurs in t 120 minutes. This rate increase with decreasing nanocluster
size demonstrates the effects of strong quantum confinement. Valence and
conduction band energy level shifts to favorable values relative to the redox
potential for the smaller 3 nm MoS2 clusters allowing the complete mineralization of PCP. This level of PCP destruction using visible light excitation of
MoS2 clusters is higher than that for TiO2 using full lamp irradiation
(300 nm , l , 700 nm) (85).
The groundwork showing the effective use of nanosized MoS2 as a photocatalyst for the destruction of toxic waste was laid by Wilcoxon et al. (89 92).
Very few other groups have studied the photocatalytic behavior of nanosized
SLTMS materials. One other group recently reported using MoS2 and WS2
nanoclusters as TiO2 sensitizers in studying the photocatalytic degradation
of 4-chlorophenol (93). Due to quantum size effects exhibited by the n-MoS2
and n-WS2, they were able to use visible irradiation for the successful destruction of the 4-chlorophenol.

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

Catalytic Properties of SLTMS

Figure 28: Visible lightinduced photooxidation of phenol as a function of time as


determined form HPLC. Order of phenol destruction: 4.5 nm MoS2 . 810 nm MoS2 . P25 TiO2
(no activity) The error bar represents reproducibility (83).

Bulk MoS2 has been studied extensively. However, nanosized MoS2 is still
in its infancy. The potential for use of this material in many aspects of
photocatalysis is very promising. The next step is to see how it performs in
the photocatalytic production of hydrogen through the splitting of water.

11.

CONCLUSIONS

The transition metal sulfide (TMS) materials have been extraordinarily useful
materials as catalysts and lubricants. They have also been the subject of
numerous fundamental and applied research programs over the past 100
years. The materials are two-dimensional with only weak van der Waals
forces holding the layers together. Thus, it has been recognized recently that
single layered transition metal sulfides (SLTMS) are generally the stable
state in heterogeneous catalytic reactions and that understanding their structure/function properties requires that we study them as single layers and in
chemically stabilized catalytic states. As single layers, this class of catalytic

35

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

36

R. R. Chianelli et al.

Figure 29: Semi-log plot of pentachlorophenol (PCP) concentration vs. visible light
illumination time. Nanosize MoS2 photocatalysts are compared to conventional
semiconductor powders: MoS2, TiO2 and CdS. The smallest MoS2 clusters (3 nm) with the most
positive oxidation potential are the most active (85).

materials will continue to be key actors in converting heavy hydrocarbon


resources and reducing pollution emission from fuels. New applications such
as photocatalytic applications will become increasingly important as the need
for environmentally neutral fuels increases. Finally, we may conclude that
important selectivity effects such as alkali promotion in alcohol synthesis
may require an understanding of the stacking of layers and intercalation of
metals. However, the long-term stability of stacked configurations in catalytic
environments is questionable.

12.

ACKNOWLEDGMENTS

We are particularly indebted to SSRL (Stanford Synchrotron Radiation


Laboratory), a national user facility operated by Stanford University on
behalf of the U.S. Department of Energy and Office of Basic Energy Sciences
under the U.S. Department of Energy Gateway Program. We would also like
to thank Professor Miguel Jose-Yacaman and the department of Chemical
Engineering and Center for Nano and Molecular Science and Technology at

Catalytic Properties of SLTMS

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

the University of Texas at Austin for the HRTEM analysis and the Robert
A. Welch Foundation for financial support. We give a special thanks to Dave
Rendina (Refinery Sciences Corp.) for encouragement and useful discussions
and to Dr. Carolina Kretschmer for revising the manuscript.

REFERENCES
[1] Weisser, O. and Landa, S. (1973) Sulfide Catalysts, Their Properties and Applications; Pergamon Press: New York.
[2] Chianelli, R.R., Daage, M., and Ledoux, M. (1994) Advances in Catalysis, 40:
177 232.
[3] Occelli, M.L. and Chianelli, R.R. (1996) Hydrotreating Technology for Pollution
Control: Catalysts, Catalysis, and Processes (Chemical Industries); Dekker/CRC
Press.
[4] Breysse, M., Furimsky, E., Kasztelan, S., Lacroix, M., and Perot, G. (2002) Catalysis Reviews, 44 (4): 651 735.
[5] Murray, H.H., Kelty, S.P., Day, C.S., and Chianelli, R.R. (1993) Inorganic Chem.
Comm., 33 (19): 4418.
[6] Schonfeld, B., Huang, J.J., and Moss, S.C. (1983) Acta Cryst., B39: 404.
[7] Wilson, J.A. and Yoffe, A.D. (1969) Adv. Phys., 18: 193.
[8] Chianelli, R.R., Prestridge, E.B., Pecoraro, T.A., and DeNeufville, J.P. (1979)
Science, 203: 1105.
[9] Tenne, R., Margulis, L., Genut, M., and Hodes, G. (1992) Nature, 360: 444.
[10] Bronson, M., Hansen, T.W., and Jacobsen, C.J.H. (2002) JACS Comm., 124: 11582.
[11] Hisashi, U., Tadao, S., and Michiko, Y. (1986) Kokai Tokkyo Koho, Japan.
[12] Chianelli, R.R. (1976) J. Crystal Growth, 34: 239.
[13] Chianelli, R.R., Scanlon, J.C., and Rao, B.M.L. (1978) J. Electro Chem. Soc.,
125 (10): 1563.
[14] Benavente, E., Santa Ana, M.A., Mendizabal, F., and Gonzalez, G. (2002) Coordination Chemistry Reviews, 224: 87.
[15] Thompson, A.H., Gamble, F.R., and Symon, C.R. (1975) Mat. Res. Bull., 10: 915.
[16] Chianelli, R.R. and Dines, M.B. (1978) Inorg. Chem., 17: 2758.
[17] Chianelli, R.R. (1982) Int. Rev. Phys. Chem., 2: 127.
[18] Gee, M.A., Frindt, R.F., Joensen, P., and Morrison, S.R. (1986) Mater. Res. Bull., 20:
543.
[19] Kosidowski, L. and Powell, A.V. (1998) Chem. Commun., 2201.
[20] Vanchura, B.A., Ho, P., Antochshuk, V., Jaroniec, M., Ferryman, A., Barbash, D.,
Fulghum, J.E., and Huang, S.D. (2002) J. Am. Chem. Soc., 124: 12090.
[21] Tauster, S.J., Percoraro, T.A., and Chianelli, R.R. (1980) J. Catal., 63 (2): 515.
[22] Bahl, O.P., Evans, E.L., and Thomas, J.M. (1968) Proc. Royal Soc. London Ser.,
A306: 53.

37

38

R. R. Chianelli et al.
[23] Chianelli, R.R., Ruppert, A.F., Behal, S.K., Kear, B.H., Wold, A., and Kershaw, R.
(1985) J. Catal., 92: 56.
[24] Tanaka, K. and Okuhara, T. (1982) The third climax conference on molybdenum
1979. J. Catal., 78: 155.
[25] Salmeron, M. and Somorjai, G.A. (1983) Surf. Sci., 127: 526.

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

[26] Farias, M.H., Gelman, A.J., Somorjai, G.A., Chianelli, R.R., and Liang, K.S. (1984)
Surf. Sci., 140: 181.
[27] Roxlo, C.B., Deckman, H.W., Gland, J., Cameron, S.D., and Chianelli, R.R. (1987)
Science, 235: 1629.
[28] Roxlo, C.B., Daage, M., Ruppert, A.F., and Chianelli, R.R. (1986) J. Catal., 100 (1):
176.
[29] Johnston, D.C., Jacobson, A.J., Silbernagel, B.G., Frysinger, S.P., Rich, S.M., and
Gebhard, L.A. (1984) J. Catal., 89 (2): 244.
[30] Johnston, D.C., Silbernagel, B.G., Daage, M., and Chianelli, R.R. (1985)
PreprintsAmerican Chemical Society, Division of Petroleum Chemistry, 30 (1):
206.
[31] Garoff, S. and Hanson, C.D. (1981) Applied Optics, 20 (5): 758.
[32] Klier, K., Beretta, A., Sun Qun, Feeley, O.C., and Herman, R.G. (1997) Catalysis
Today, 36: 3.
[33] Chang, C.H. and Chan, S.S. (1981) J. Catal., 72 (1): 139.
[34] Berhault, G., Mehta, A., Pavel, A., Yang, J., Rendon, L., Jose-Yacaman, M., CotaAraiza, L., Moller, A., and Chianelli, R.R. (2001) J. Catal., 198: 9.
[35] Helveg, S., Lauritsen, J.V., Laegsgaard, E., Stensgaard, I., Nrskov, J.,
Clausen, B.S., Topse, H., and Besenbacher, F. (2000) Phys. Rev. Letters, 84 (5): 851.
[36] Silbernagel, B.G., Pecoraro, T.A., and Chianelli, R.R. (1982) J. Catal., 28 (2): 380.
[37] Liang, K.S., Hughes, G.J., and Chianelli, R.R. (1984) J. Vac. Sci. Technol., A2 (2):
991.
[38] Carlsson, A., Brorson, M., and Topse, H.J. (2004) J. Catal., 227: 530.
[39] Spirko, J.A., Neiman, M.L., Oelker, A.M., and Klier, K. (2003) Surface Science, 542:
192.
[40] Spirko, J.A., Neiman, M.L., Oelker, A.M., and Klier, K. (2004) Surface Science, 572:
191.
[41] Lauritsen, J.V., Nyberg, M., Nrskov, J.K., Clausen, B.S., Topse, H.,
Laesgaard, E., and Besenbacher, F. (2004) J. Catal., 224: 94.
[42] Daage, M. and Chianelli, R.R. (1994) J. Catal., 149: 414.
[43] Eijsbouts, S. and VanLeerdam, G.C. (1995) Bull. Soc. Chim. Belg., 104: 347.
[44] Shido, T. and Prins, R. (1998) J. Phys. Chem., B102: 8426.
[45] Perez De la Rosa, M., Texier, S., Berhault, G., Camacho, A., Jose-Yacaman, M.,
Mehta, A., Fuentes, S., Montoya, J.A., Murrieta, F., and Chianelli, R.R. (2004)
J. Catal., 225: 288.
[46] Liang, K.S., Chianelli, R.R., Chien, F.Z., and Moss, S.C. (1986) J. Non-Cryst.
Solids, 79: 251.

Catalytic Properties of SLTMS


[47] Peng, Y., Meng, Z., Zhong, C., Lu, J., Yu, W., Yang, Z., and Qian, Y. (2001) J. Solid
State Chem., 159: 170.
[48] Glasson, C., Geantet, C., Lacroix, M., Labruye`re, F., and Dufresne, P. (2002)
J. Catal., 212: 76.

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

[49] Sherwood, P.W. (2002) High pressure hydrogenation in Germany. I: The Liquid
PhaseU.S. Department of Commerce, Technical Final Industrial Intelligence
Division:p.2; FAIT final Report No. 952.
[50] Zorn, H., inventor: I.G. Farbenindustrie, assignee (1932) Conversion of higher
boiling hydrocarbons to those of lower boiling point. U.S. Patent 1,876,270.
[51] Kretschmar, K., Merz, L., Niemana, K., Guitian, J., Krasuk, J., and Marruffo, F.,
inventors; Veba OEL and Intevep S.A., assignees. (1989) Process for the hydrogenation of heavy residual oils. U.S. Patent 4,851,107.
[52] Braun, R.; Johnson, R.W.; Gatsis, J.G., inventors; UOP, Inc., assignee. (1981).
Catalytic slurry process for black oil conversion. U.S. Patent 4,285,80.
[53] Ranganathan, R.; Denis; Jean-Marie, D.; Pruden, B.B., inventors; Energy Mines
and Resources Canada, assignee, Hydrocracking of heavy oils using iron coal
catalyst. U.S. Patent 4,214,977 (1980).
[54] Benham, N.K. and et al. (1996) Canmet Residuum Hydrocracking Advances
through Control of Polar Aromatics. Paper AM-96-58, NPRA Annual Meeting,
San Antonio, Texas, March 1719.
[55] Krasuk, J.H.; Silva, F.J.; Galiasso, R.E.; Souto, A., inventors: Intevep, S.A.,
assignee. (1986) Process for hydroconversion and upgrading of heavy crudes of
high metal and asphaltene content. U.S. Patent 4,591,426.
[56] Bearden, R., Jr. and Aldridge, C.L. inventors: Exxon Research and Engineering
Company, assignee. (1979) Hydroconversion of heavy hydrocarbons. U.S. Patent
4,134,825.
[57] Bearden, R., Jr. and Aldridge, C.L. (1981) Energy Progress, 1 (14): 44. Note that
M-Coke (micrometallic coke) was used to designate catalyst in this publication.
M-coke was subsequently dropped in favor of Microcat.
[58] Silverman, M.A. (1995) (Stone and Webster), SOC Technology-A Flexible Approach
to Residual Oil Upgrading, AIChE Spring Meeting, Houston, Texas, March 19 23.
[59] U.S. Patent 5,578,197, 1996.
[60] Montanari, R., Dellagiovanna, S., Rosi, S., Panoriti, N., Marchionna, M., and Del
Bianco, A. (2003) Convert Heaviest Crude and Bitumen into Extra Clean
Products, 2003 NPRA Annual Meeting, San Antonio, Texas, March 2325.
[61] Panariti, N., Del Bianco, A., Del Piero, G., and Marchionna, M. (2000) Appl. Catal.
A. Gen., 204: 203.
[62] Kramer, David C.; inventor; Chevron Research and Technology Company, assignee.
(1994) Process to prevent catalyst deactivation in activated slurry processing. U.S.
Patent 5,298,152.
[63] Peries, J.P., Raimbault, C., des Courieres, J., and Gouzien, L. (1990) Tervahl
Process at the Border Between Thermal and Catalytic Upgrading Processes,
1990 NPRA Annual Meeting, March 2527.
[64] Giuliani, P., Le Page, Jean-Francois; Plumail, Jean-Claude; Espaillac, M.,
inventors. Institute Francais du Petrole, assignee. (1991) Catalytic composition
comprising a metal sulfide suspended in a liquid containing asphaltenes and
hydrovisbreaking process of a hydrocarbon charge. U.S. Patent 5,024,751.

39

40

R. R. Chianelli et al.

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

[65] Poole, M.C., Bauman, R.F., Bearden, R., Jr., Effron, E., and Wu, F.S. (1994)
Microcat-RCw Upgrading Option for Heavy Crudes and Residua. AIChE Annual
Meeting, Atlanta, Georgia, April 21.
[66] Transmission electron micrographs provided by G.W. Bailey of Exxon Research and
Development Labs in 1978. The micrographs were taken on a Phillips EM 300
instrument. Toluene washed, dry catalyst powders were dispersed in isopropyl
alcohol using sonocation. A drop of dispersion was then placed on a carbon
coated copper grid and alcohol was evaporated. Micrographs were taken at high
magnification and then further enlarged photographically to provide the final
magnified images.
[67] Bearden, R., Jr. (1997) Microcat-RC: Technology for the Hydroconversion
Upgrading of Petroleum Residua. Presentation for the George A. Olah Award in
hydrocarbon or petroleum chemistry, Petroleum Division, 213th ACS National
Meeting, San Francisco, California, April 13 17.
[68] Bell, R.E. and Herfert, R.E. (1957) J. Am. Chem. Soc., 79 (3): 351.
[69] Bearden, R., Aldridge, C. L.; inventors; Exxon Research and Engineering Company,
assignee. (1987) Hydrocracking with phosphomolybdic acid and phosphoric acid.
U.S. Patent 4,637,870.
[70] Bearden, R., Aldridge, C. L., inventors; Exxon Research and Engineering Company,
assignee. (1991) Hydroconversion process using a sulfided molybdenum catalyst
concentrate. U.S. Patent 5,039,392.
[71] Killerfer, D.H. and Linz, A. (1952) Molybdenum Compounds; Interscience Publishers: New York, NY, 88.
[72] Conradson Carbon was measured using ASTM Test D-189.
[73] Schucker, R.C. (1983) Ind. Eng. Chem. Process Des., Dev., 22 (4).
[74] Aldridge, C. L., Bearden, R., inventors. Exxon Research and Engineering Company,
assignee. (1981) High Surface Area Catalysts. U.S. Patent 4,244,839.
[75] Micrograph provided by G.W. Bailey and Ezra Shannon of Exxon Research and
development Laboratories in 1981. See footnote 66 for procedure.
[76] U.S. Patent 4,244,839, 1981.
[77] Moll, N.G. and Quarderer, G.J. (1979) Chem. Engin. Prog., 75 (11): 46.
[78] Aldridge, C. L., Bearden, R., inventors; Exxon Research and Engineering Company,
assignee. (1983) Coal Liquefaction Process. U.S. Patent 4,369,106.
[79] Aldridge, C. L., Bearden, R., inventors; Exxon Research and Engineering Company,
assignee. (1981) Hydroconversion of an oil-coal mixture. U.S. Patent 4,298,454.
[80] Application of exfoliation techniques to the preparation of MoS2 liquefaction
catalysts. Bockrath, Bradley, C. and Parfitt, D.S. (1995) United States Department
Energy, Pittsburgh Energy Technology Center: Pittsburgh, PA, USA. Coal Sci.
Technol. 24(Coal Science, Vol. 2), 13431346.
[81] FETC Publications (1997) Conference Proceedings, Coal Liquefaction & Solid Fuels
Contractors Review Conference.
[82] Ahmed, S.M. (1982) Electrochim. Acta, 27 (6): 707.
[83] Thurston, T.R. and Wilcoxon, J.P. (1999) J. Phys. Chem., B, 103 (1): 11.
[84] Wilcoxon, J.P. and Thurston, T.R. (1999) Materials Research Society Symposium
Proceedings, Symposium on Advanced Catalytic Materials-1998 at the 1998 MRS
Fall Meeting, 549: 119.

Catalytic Properties of SLTMS


[85] Wilcoxon, J.P. (2000) J. Phys. Chem. B, 104 (31): 7334.
[86] Mills, G. and Hoffmann, M.R. (1993) Environ. Sci. Technol., 27 (8): 1681.
[87] Hoffmann, M.R., Martin, S.T., Choi, W.Y., and Bahnemann, D.W. (1995) Chem.
Rev., 95 (1): 69.
[88] Serpone, N., Maruthamuthu, P., Pichat, P., Pelizzetti, E., and Hidaka, H. (1995)
J. Photochem. Photobiol. AChem., 85 (3): 247.

Downloaded by [Instituto Mexicano del Petroleo] at 11:31 02 May 2016

[89] Wilcoxon, J.P., Newcomer, P.P., and Samara, G.A. (1997) J. Appl. Phys., 81 (12):
7934.
[90] Wilcoxon, J.P., Martino, A., Baughmann, R.L., Klavetter, E., and Sylwester, A.P.
(1993) Materials Research Society Conference Proceedings, Symp. on Nanophase
and Nanocomposite Materials, Komarneni, S., Parker, J. and Thomas, G., eds.;
Materials Research Soc: Boston, MA, Vol. 286, 131.
[91] Wilcoxon, J.P., Martino, A., Baughmann, R.L., Klavetter, E., and Sylwester, A.P.
(1994) Advanced Study Institute on Nanophase Materials Series E, Applied
Sciences: SynthesisPropertiesApplications, Hadjipanayis, G. and Siegel, R.,
eds.; Kluwer Academic Publishers: Corfu, Greece, Vol. 260, 780.
[92] Wilcoxon, J.P. and Samara, G.A. (1995) Phys. Rev. B (Condensed Matter), 51 (11):
7299.
[93] Ho, W.K., Yu, J.C., Lin, J., Yu, J.G., and Li, P.S. (2004) Langmuir, 20 (14): 5865.

41

Você também pode gostar