Sie sind auf Seite 1von 8

Journal of Catalysis 317 (2014) 2229

Contents lists available at ScienceDirect

Journal of Catalysis
journal homepage: www.elsevier.com/locate/jcat

Kinetics and mechanism of m-cresol hydrodeoxygenation


on a Pt/SiO2 catalyst
Lei Nie, Daniel E. Resasco
School of Chemical, Biological, and Materials Engineering, The University of Oklahoma, Norman, OK 73019, USA

a r t i c l e i n f o a b s t r a c t

Article history: The LangmuirHinshelwood kinetic model was employed to investigate the hydrodeoxygenation (HDO)
Received 2 February 2014 of m-cresol in the vapor phase over Pt/SiO2, at 300 C. The model takes into account the competitive
Revised 26 May 2014 adsorption and secondary reactions of all possible reactants and products. Therefore, independent
Accepted 29 May 2014
measurements of the conversion of the reaction intermediates 3-methyl-cyclohexanone and 3-methyl-
Available online 2 July 2014
cyclohexanol were done at various levels of conversion and incorporated in the kinetic tting of the
model. The kinetic analysis of these reactions over the Pt/SiO2 catalyst conrms that hydrogenation
Keywords:
and dehydrogenation reactions occur at faster rates than the hydrodeoxygenation steps. The hydrogena-
Lignin
Phenolic compound
tion/dehydration followed by dehydrogenation (HYD route) was shown to be a minor deoxygenation
Cresol pathway. The actual hydrodeoxygenation path is proposed to proceed via a tautomerization route that
Pt/SiO2 starts with a fast and reversible tautomerization of m-cresol to an unstable ketone intermediate (3-
Hydrodeoxygenation methyl-3,5-cyclohexadienone) over the catalyst surface. In the case of the Pt/SiO2 catalyst, the C@C bonds
Hydrogenation of the unstable ketone intermediate are fully hydrogenated to the saturated ketone (3-methyl-cyclohex-
Tautomerization anone). Sequential hydrogenation of this ketone produces a saturated alcohol (3-methyl-cyclohexanol).
In the absence of surface acidity, this path only leads to the two ring-saturated products. When the
support is acidic enough, dehydration of the alcohol to 3-methyl-1-cyclohexene followed by dehydroge-
nation leads to toluene. The kinetic analysis conducted here shows that this path is very limited over Pt/
SiO2. By contrast, the toluene selectivity can be dramatically enhanced when using TiO2 and ZrO2 as the
catalyst supports. This enhancement is not through the hydrogenation/dehydration/dehydrogenation
path, but rather through the direct dehydroxylation mentioned above, which is favored by the incorpo-
ration of oxophilic sites in the catalysts. The oxophilic sites increase the rate of chemoselective hydroge-
nation of the carbonyl group of the m-cresol tautomer (3-methyl-3,5-cyclohexadienone) to a very
reactive unsaturated alcohol (3-methyl-3,5-cyclohexadienol), which can be easily dehydrated to toluene,
even in the absence of any signicant catalyst acidity.
2014 Elsevier Inc. All rights reserved.

1. Introduction guaiacols, and syringols) represent an important fraction [12,13].


The two main oxygenated groups that appear in these compounds
The liquid product from the fast pyrolysis of biomass (i.e., attached to the aromatic ring are the hydroxy (ArAOH) and the
bio-oil) requires a signicant degree of chemical transformation methoxy (ArAOACH3) groups. In order to gain a detailed under-
to produce transportation fuels. Oxygen removal is one of the most standing of the fundamental chemistry beyond the upgrading of
important upgrading reactions since the high oxygen content is phenolics, researchers have conducted extensive studies using
responsible for most of the undesirable properties of bio-oil, model phenolic compounds such as phenol, cresols, anisole, and
including corrosivity, low thermal stability, low volatility, and guaiacol [1417].
low energy content [14]. Over the past few decades, extensive Elucidating the detailed reaction mechanism of the hydrodeox-
research and development efforts have focused on the hydropro- ygenation (HDO) of phenolics has been the goal of a number of
cessing of bio-oil [511]. Among the hundreds of compounds recent studies [1523]. Yet, as discussed in this contribution, there
present in bio-oil, the lignin-derived phenolics (phenols, anisoles, are aspects of the mechanism that remain unsettled and require
further analysis. Most of the proposed deoxygenation mechanisms
reported in the literature can be grouped into two major pathways,
Corresponding author. Fax: +1 405 325 5813. (a) primary hydrogenation/secondary deoxygenation (HYD) route
E-mail address: resasco@ou.edu (D.E. Resasco). [15,16,2427], and (b) direct deoxygenation (DDO) route via CAO

http://dx.doi.org/10.1016/j.jcat.2014.05.024
0021-9517/ 2014 Elsevier Inc. All rights reserved.
L. Nie, D.E. Resasco / Journal of Catalysis 317 (2014) 2229 23

hydrogenolysis [22,23,28,29]. The HYD route typically starts with support is inert enough to minimize the dehydration activity. Fur-
hydrogenation of the aromatic ring of phenolic compounds to their thermore, we have selected a moderate reaction temperature, at
corresponding alcohols. Subsequently, oxygen removal could occur which the toluene/methylcyclohexane equilibrium ratio is low
via CAOH hydrogenolysis or dehydration [15,18]. Usually, the HYD enough that has allowed us to follow the evolution of these two
route has been proposed to occur over bifunctional catalysts with a crucial hydrocarbon products as a function of conversion toward
metal function that hydrogenates the aromatic ring and an acid equilibrium, giving us further insight into the mechanism. In
function that catalyzes the removal of the O atom by dehydration addition, the measurements of the m-cresol conversion rate were
[20]. For example, in a recent study [16], the hydrodeoxygenation combined with direct measurement of the conversion rates of each
of phenol was investigated over a bifunctional Ni/HZSM-5 catalyst individual intermediate product, which provides further validation
in a batch reactor at 5 MPa H2 and 200 C. It was proposed that a to the kinetic side reaction constants, and the thermodynamic
partial hydrogenation of the aromatic ring takes place over the adsorption constants. These consistent kinetics data have been
Ni metallic function to form cyclohexanone (via tautomerization) analyzed by using a LangmuirHinshelwood model, which results
and cyclohexanol, followed by dehydration of the alcohol over in meaningful thermodynamic and kinetics parameters. They can
the acid sites provided by HZSM-5 to yield cyclohexene. Finally, be used to obtain a more detailed picture of the HDO mechanism
cyclohexene was hydrogenated to cyclohexane, which was the than those previously proposed.
major hydrocarbon observed at this temperature [14,16].
The direct deoxygenation (DDO), i.e., the direct cleavage of the
2. Experimental
Csp2AO bond via hydrogenolysis, has been proposed as a hydro-
deoxygenation pathway for phenolic compounds on CoMo sulde
2.1. Catalyst preparation and characterization
catalysts. For example, Massoth et al. [19,30] have proposed that
the DDO route dominates over the HYD route when converting
As summarized in Table 1, the catalyst used in this study was a
phenol and cresol over suldes at high temperatures. Based on
1 wt.% Pt/SiO2 prepared by incipient wetness impregnation of the
the observation that benzene (and toluene) appeared as primary
SiO2 support (HiSil 210, PPG, SBET = 135 m2/g) with an aqueous
products, it was concluded that DDO Csp2AO hydrogenolysis
solution of H2PtCl66H2O (Aldrich), dried overnight at room tem-
mainly takes place at vacancy sites that expose Mo (or Co), while
perature and then at 110 C for 12 h, followed by calcination at
the adjacent HYD sites are S or SH saturated sites. As will be
400 C for 4 h. To estimate the metal dispersion, the CO uptake
discussed below, the high strength of the Csp2AO bond makes
was measured at room temperature in a dynamic pulse unit, giving
CAO hydrogenolysis a very difcult step that could only take place
a value of CO/Pt = 0.28. The catalyst was further characterized by
at very high temperatures due to the high energy barriers required
high-resolution electron microscopy. Fig. 1 shows an HRTEM
by the cleavage of this strong bond.
image of the catalyst after reduction at 300 C and passivation at
In a recent study conducted at mild temperatures over a
room temperature. The average particle size estimated by counting
bimetallic catalyst on a nonacidic support [31], we proposed that
100 metal particles is about 3.5 nm, in good agreement with the
a fast enol-keto tautomerization of m-cresol step precedes the ring
chemisorption results. The pre-reduction in the catalyst was con-
hydrogenation. In that case, an unstable ketone intermediate
ducted in owing H2 at 300 C at atmospheric pressure, before each
(3-methyl-3,5-cyclohexadienone) would form over the catalyst
run or characterization analysis, as detailed in previous work [12].
surface, which could be hydrogenated via two possible paths.

(a) When the carbonyl group of this intermediate gets hydroge- 2.2. Reaction kinetics
nated instead of the ring, a very reactive unsaturated alcohol
(3-methyl-3,5-cyclohexadienol) is formed and it can be The vapor-phase reaction of m-cresol over Pt/SiO2 was investi-
readily dehydrated to toluene. This easy dehydration step gated on an isothermal tubular reactor operating under H2 ow at
is driven by the aromaticity of the product. As a result, this atmospheric pressure. The data were analyzed assuming an
dehydration step might even occur thermally and would isothermal plug ow reactor (PFR) operating in two modes, differ-
not require any catalyst acidity. ential (conversion <10%) and integral (entire W/F range). The pel-
(b) When the hydrogenation occurs at the C@C bonds of the letized catalyst (size range: 250425 lm) was placed at the
unstable intermediate 3-methyl-3,5-cyclohexadienone, the center of a vertical tubular quartz reactor between two layers of
typically observed 3-methylcyclohexanone is formed and is glass beads and quartz wool. Following the criteria proposed by
further hydrogenated to 3-methylcyclohexanol. Over sup- Madon and Boudart [32], calculations were done to ensure that
ports that contain acid sites strong enough to dehydrate this external and internal mass transfer limitations were eliminated.
alcohol, methylcyclohexene could be formed and a sub- The catalyst was pre-reduced in ow of H2 (120 ml/min, Airgas,
sequent dehydrogenation would produce toluene. However, 99.99%) for 1 h at 300 C. After reduction, the reaction was con-
on most inert supports dehydration of saturated alcohols ducted at the same H2 ow rate and temperature. A 0.5 ml/h ow
should not occur to a great extent. of liquid m-cresol (Aldrich, 99.5%) was fed continuously from a syr-
inge pump (Cole Parmer) and vaporized into a gas stream of
This mechanism can explain several important experimental 120 ml/min H2. The H2/feed molar ratio was kept at 7080:1 to
results that none of the previously proposed mechanisms can. minimize catalyst deactivation. The reaction products were ana-
For example, it explains how toluene can be obtained at mild tem- lyzed online by gas chromatography (Agilent model 6890) using
peratures and in the absence of strong acid sites. It also explains an HP-Innowax capillary column and a FID detector. The carbon
why toluene is formed as a primary product, before methylcyclo-
hexane, and why methylcyclohexene is not observed on this
Table 1
catalyst, even at low conversions. Characteristics of the Pt/SiO2 catalyst used for the kinetics measurements.
In this contribution, we attempt to investigate whether the pro-
Pt metal Silica support
posed tautomerization/secondary chemoselective hydrogenation
mechanism derived from a study on NiFe catalysts is also opera- Metal loading 1% Type HiSil 210
CO/Pt 0.28 Manufacturer PPG Industries
tive on a conventional catalyst as common as Pt/SiO2, in which the
dp (TEM) 3.5 nm BET 135 m2/g
simple metallic function is represented by Pt, while the SiO2
24 L. Nie, D.E. Resasco / Journal of Catalysis 317 (2014) 2229

(a) 25
Conversion R2=0.98
20 ONE
OL
TOL

Conversion/Yield (%)
15 MCH

10

0
0 2 4 6 8 10 12 14
10 nm
W/F (gcat.hr/mol)

Fig. 1. HRTEM of 1 wt% Pt/SiO2 after calcination at 400 C in air, reduction at 300 C
and passivation at room temperature.
(b) 60 R2=0.99
Conversion
50
balance was checked in every run, and it was found to be better
m-Cr
OL

Conversion/Yield (%)
than 95% in each case. 40
A rate expression based on the conventional LangmuirHinshel-
TOL
wood model that includes the partial pressures of m-cresol,
MCH
30
3-methyl-cyclohexanone, 3-methyl-cyclohexanol, toluene, methyl-
cyclohexane, and 3-methyl-1-cyclohexene was used to t the 20
experimental results. Nonlinear least-square regression analysis
of the differences between the experimental values and the kinetic 10
model calculated values was used to t simultaneously the exper-
imental data for different reactant feeds. That is, the tting of the 0
rate expression was obtained by alternatively using as feed 0 1 2 3 4
not only m-cresol, but also the intermediate compounds that W/F (gcat.hr/mol)
participate in the reaction pathway (i.e., m-cresol, 3-methyl-
cyclohexanone, 3-methyl-cyclohexanol, toluene, and methylcyclo- (c) 80
hexane) as a function of W/F. Several other constraints were used R2=0.98
to ensure that the tting resulted in the global minimum and 70 Conversion
values with reasonable physical meanings. For example, the values m-Cr
60
for the initial guesses in the tting routine were obtained from the ONE
TOL
Conversion/Yield (%)

kinetics and thermodynamic parameters calculated from the dif- 50


ferential reactor data for each compound, which eliminates the MCH
40
inuence of competitive adsorption of the products on the
measured rates. Also, as an external validation, a mixture of m- 30
cresol and 3-methyl-cyclohexanol was used as the reactant and
the evolution of partial pressures with W/F was compared to that 20
calculated using the rate expression with the parameters obtained
10
with the pure feeds.
0
0 1 2 3 4 5
3. Results and discussion
W/F (gcat.hr/mol)
3.1. Experimental measurements in integral reactor mode
Fig. 2. Conversion and yield of products of different feeds over Pt/SiO2 as a function
of W/F at 300 C. Carrier gas: H2. Pressure = 1 atm. (a) m-Cresol; (b) 3-methyl-
Fig. 2 shows the evolution of products as a function of W/F cyclohexanone; (c) 3-methyl-cyclohexanol. MCH: methyl-cyclohexane; TOL: tolu-
when the feeds were (a) m-cresol, (b) 3-methyl-cyclohexanone, ene; m-Cr: m-cresol; ONE: 3-methyl-cyclohexanone; OL: 3-methyl-cyclohexanol.
and (c) 3-methyl-cyclohexanol, respectively over Pt/SiO2 at The points are experimental data and the lines are calculated from the Langmuir
Hinshelwood kinetics model.
300 C. The symbols represent the actual experimental data,
whereas the solid lines represent the calculated values from a sin-
gle LangmuirHinshelwood model used for all feeds. This will be Fig. 1b and c, for the other two feeds, the toluene yields were less
discussed below in more detail. According to Fig. 2a, the major than 1% in the entire W/F range. This clearly indicates that, on this
products from m-cresol are 3-methyl-cyclohexanone, 3-methyl- non-acidic catalyst, neither 3-methyl-cyclohexanone nor 3-
cyclohexanol, and toluene. The yield of methyl-cyclohexane is less methyl-cyclohexanol is a major precursor for the formation of
than 0.3% for all W/F values. A critical result that will be relevant in toluene.
the discussion below is that the yield of toluene from m-cresol is Fig. 2b shows the product distribution for 3-methyl-cyclohexa-
signicant, even at low conversions. In fact, the slope of the yield none conversion over Pt/SiO2 as a function of W/F. m-Cresol and 3-
curve when W/F approaches 0 is not zero, which suggests that tol- methyl-cyclohexanol are the two major products, while toluene
uene may be a primary product. By contrast, as shown in both and methyl-cyclohexane yields are both lower than 1%. Similarly,
L. Nie, D.E. Resasco / Journal of Catalysis 317 (2014) 2229 25

Equilibrium*
C: 33%

OH

kDOH

kCK kCO
kOC kET kTM
O kKC
OH
kKO kDW
kOK kEM
Equilibrium* Equilibrium*
K : 52% O : 15%

Scheme 1. Reaction pathway over Pt/SiO2 with rate constant. (Denotation) kCK: rate constant for m-cresol to 3-methyl-cyclohexanone. (Subscript) C: m-cresol; K: 3-methyl-
cyclohexanone; O: 3-methyl -cyclohexanol; E: 3-methyl-1-cyclohexene; T: toluene; M: methyl-cyclohexane; DOH: dehydroxylation; DW: dehydration. : equilibrium molar
% between m-cresol, 3-methyl-cyclohexanone and 3-methyl-cyclohexanol under this condition.

Table 2
Product distribution, rate and rate constant for 3-methyl-1-cyclohexene over Pt/SiO2 and 3-methyl-cyclohexanol over SiO2 at 300 C. Pressure = 1 atm. MCH: methyl-
cyclohexane; TOL: toluene; OL: 3-methyl-cyclohexanol; 3-Me-CYHE: 3-methyl-1-cyclohexene; Me-CYHE: methyl-cyclohexene.

Feed Catalyst W/F ((gcat h)/mol) Product yield (%) rDW rET rEM kDW kET kEM
TOL MCH Me-CYHE Isomers mol/(gcat h)
3-Me-1-CYHE Pt/SiO2 0.06 1.7 1.9 9.0 0.28 0.32 1.7 2.0
0.10 4.7 3.9 10.3
OL SiO2 12.5 1.4 0.002 0.005

Fig. 2c shows the product distribution for 3-methyl-cyclohexanol when the HDO route is operative, the intermediate olen is usually
conversion over Pt/SiO2 as a function of W/F. In this case, again, tol- not detected in the products or at very low amounts due to its high
uene and methyl-cyclohexane yields are both lower than 1%, while activity. At the low W/F values measured here, in addition to
m-cresol and 3-methyl-cyclohexanone are the two major products. toluene and methyl-cyclohexane, the isomerization products
Even from a qualitative observation of the data in Fig. 2, it can 1-methyl-1-cyclohexene and 4-methyl-1-cyclohexene were
be concluded that m-cresol, 3-methyl-cyclohexanone, and obtained from 3-methyl-1-cyclohexene.
3-methyl-cyclohexanol interconvert at faster rates than the In addition to the reversible hydrogenation/dehydrogenation
hydrodeoxygenation to hydrocarbons. However, while the hydro- steps of the oxygenated compounds, Scheme 1 includes a direct
genation and de-hydrogenation rates between these three compo- hydrodeoxygenation pathway from m-cresol to toluene, with rate
nents are relatively high, they are still far from equilibrium at all constant denoted as kDOH. As mentioned above and further dis-
W/F values investigated. The equilibrium composition at 300 C cussed below, this direct dehydroxylation step does not involve a
for m-cresol, 3-methyl-cyclohexanone and 3-methyl-cyclohexanol hydrogenolysis step, but rather a tautomerization/secondary
is 33:52:15 in molar percentage, obtained from our kinetic chemoselective hydrogenation path. The dehydration step of 3-
analysis. Therefore, as illustrated in Scheme 1, while m-cresol, methyl-cyclohexanol to 3-methyl-1-cyclohexene (minor in this
3-methyl-cyclohexanone, and 3-methyl-cyclohexanol interconvert case) is denoted as kDW (dehydration). The dehydrogenation of
relatively fast, toluene is only obtained as a primary product from methyl-cyclohexane to toluene was neglected in this scheme due
m-cresol, which indicates the participation of a direct deoxygen- to the very low yields of methyl-cyclohexane observed in all cases.
ation reaction pathway. The LangmuirHinshelwood model that follows is based on this
In a study of m-cresol conversion over Pt/c-Al2O3 at 260 C, reaction scheme and proved to represent very well the experimen-
Foster et al. [20] proposed that, following hydrogenation of the aro- tal behavior.
matic ring, dehydration of 3-methyl-cyclohexanol produces
methyl-cyclohexene as an intermediate, which can be sequentially
hydrogenated or dehydrogenated to methyl-cyclohexane or 3.2. Kinetic model
toluene, which is the HDO route mentioned above. However, this
cannot be a possible path over the Pt/SiO2 catalyst investigated Previous kinetic studies of the HDO of m-cresol over sulded
here since, as shown Table 2, the dehydration activity of the SiO2 CoMo catalysts have used the LangmuirHinshelwood model
support is very low. Furthermore, this is in agreement with the [19,22,33]. In those studies, H2 was assumed to adsorb noncompet-
low toluene yield obtained when 3-methyl-cyclohexanol was used itively over sites different from those used by the phenolic reactant
as a feed on Pt/SiO2, as shown in Fig. 2c. and intermediates. In all cases, the authors considered two parallel
In fact, if dehydration occurred, the olen thus produced would reaction pathways: (1) a direct dehydroxylation via hydrogenolysis
be rapidly converted. As also shown in Table 2, when 3-methyl-1- of the ArAOH bond (direct deoxygenation, DDO route) leading to
cyclohexene was fed over Pt/SiO2 at 300 C, the conversion was toluene formation (which, should not occur under the milder con-
high, even at very low space times. For example, at W/F = 0.06 ditions or our study) and (2) a pre-hydrogenation of the aromatic
0.10 (gcat h)/mol, conversions of 12.6% and 18.9% were achieved, ring to the corresponding alcohol followed by dehydration to
which show the very fast reaction rates for this molecule. This cyclohexene and subsequent hydrogenation to cyclohexane (HYD
agrees well with literature reports [15,16,18,20], which show that route).
26 L. Nie, D.E. Resasco / Journal of Catalysis 317 (2014) 2229

In this contribution, since the reaction temperature is mild, (iii) Hydrogen adsorbs dissociatively on the metal [34], but it
deoxygenation through direct ArAOH hydrogenolysis is considered does not compete for the same sites with the phenolic
negligible. This will be further discussed in the discussion part. molecules. It is well known that the small H atom preferen-
However, a direct deoxygenation pathway is still included in the tially adsorbs on high-coordination 3-fold and 4-fold hollow
reaction scheme. In the discussion part of this contribution, it will sites on a metal surface [35,36]. Therefore, in hydrogenation/
be shown that instead of direct hydrogenolysis of the ArAOH bond, dehydrogenation reactions of organic compounds, the larger
this direct dehydroxylation is proposed to occur via the tautomer- molecules occupy one or more top sites without being inhib-
ization/secondary hydrogenation pathway. ited by an adsorbed H atom at a lower hollow site. As a result,

0
p !
kDOH  K c  PC K H2  P H2
r DOH  p 1
1 K c  Pc K ONE  PONE K OL  POL K TOL  PTOL K MCH  PMCH K ENE  PENE 1 K H2  P H 2

The effect of co-feeding water with the reactant is shown in the co-adsorption of H can be considered as an independent
Fig. 3. It is observed that the conversion of m-cresol as well as adsorption on different sites (S1 and S2) as summarized in the
the product distribution is not remarkably affected by the presence sequence of elementary steps included in Fig. 4 [37].
of water, even when the partial pressure of the water was six times (iv) Surface reaction is rate-determining step.
higher than that of m-cresol. This result supports the assumption
that the adsorption of water has little impact on the reaction rate From these assumptions, the rate expression for the deoxygen-
and that dehydration does not play a signicant role, but rather ation of m-cresol to toluene can be written as follows:
the activity of the metal dominates the kinetics. Since the partial pressure of hydrogen is in excess of the amount
Based on this qualitative analysis, a rate expression was derived consumed and remains practically constant across the reactor, the
0
based on the conventional LangmuirHinshelwood model, with last term in Eq. (1) is a constant that can be lumped into the kHDO
the following assumptions: constant, giving:

kDOH  K c  PC
r DOH 2
1 K c  Pc K ONE  PONE K OL  POL K TOL  P TOL K MCH  PMCH K ENE  PENE

(i) Competitive molecular adsorption of m-cresol, 3-methyl- where


cyclohexanone, 3-methyl-cyclohexanol, toluene, methyl- p !
cyclohexane, and 3-methyl-1-cyclohexene occurs over the 0 K H2  P H2
kDOH kDOH  2:1
same metal sites on the catalyst. 1 K H2  P H 2
(ii) Rate inhibition by water adsorption is neglected.

3.3. Experimental measurements in differential reactor mode

To obtain meaningful tting parameters, we have made inde-


25 pendent measurements in the differential reactor (low conversion),
which allows us to simplify the rate expression by ignoring the
adsorption terms for the products and obtaining the reactant
20 Conversion parameters. Thus, under differential conditions for any reactant j
Conversion/Yield (%)

MCH
TOL kj  K j  P j
ONE rj 5
15 1 K j  Pj
OL
which can be rearranged as the following linear expression,
10  
1 1 1 1
 5:1
rj kj K j P j kj
5
By plotting 1/rj vs. 1/Pj, the adsorption constant Kj of any given
 
reactant can be obtained from the ratio of the intercept k1j to the
0  
0 1 2 3 4 5 6 slope kj1K j [37]. Here, again, the H2 term is lumped into the rate
Pwater / Pm-Cr constant k and does not affect the calculation of Kj.
Independent measurements at varying reactant partial
Fig. 3. Effect of co-feeding water on the product distribution for m-cresol pressures were made under differential conditions for m-cresol,
conversion over Pt/SiO2 as a function of partial pressure ratio between water and
m-cresol. Temperature = 300 C. Pressure = 1 atm. Pm-Cr = 1486 Pa. MCH: methyl-
3-methyl-cyclohexanone (ONE), 3-methyl-cyclohexanol (OL), tolu-
cyclohexane; TOL: toluene; m-Cr: m-cresol; ONE: 3-methyl-cyclohexanone; OL: 3- ene (TOL), methyl-cyclohexane (MCH) and 3-methyl-1-cyclohex-
methyl-cyclohexanol. ene (ENE). Each of them was run independently as a reactant
L. Nie, D.E. Resasco / Journal of Catalysis 317 (2014) 2229 27

KC K TOL
Cr + S1 Cr - S1 TOL + S1 TOL - S1
K ONE K MCH
ONE + S1 ONE - S1 MCH + S1 MCH - S1
K OL K ENE
OL + S1 OL - S1 ENE + S1 ENE - S1

k DOH K H2
Cr - S1 + 2H 2 TOL - S1 + H2O + 2S2
H2 + 2S2 2H 2
k TM
TOL - S1 + 6H 2 MCH - S1 + 6S2
k CK Cr: m-cresol;
Cr - S1 + 4H ONE - S1 + 4S2
2 TauCr: m-cresol tautomer
k KC (3-methyl-3,5-cyclohexadienone);
ONE - S1 + 4S2 Cr - S1 + 4H 2
ENE: 3-methyl-1-cyclohexnene;
k KO ONE: 3-methyl-cyclohexanone;
ONE - S1 + 2H-S2 OL - S1 + 2S2 OL: 3-methyl-cyclohexanol;
k OK
TOL: toluene;
k OC
OL - S1 + 6S2 Cr - S1 + 6H-S2 MCH: methyl-cyclohexane;
k CO S1: active site type 1;
k DW S2: active site type 2.
OL - S1 ENE - S1 + H2O-S1
k ET
ENE - S1 TOL - S1 + 4H-S2
k EM
ENE - S1 + 2H-S2 MCH - S1 + 2S2

Fig. 4. Sequence of elementary steps for m-cresol conversion over Pt/SiO2.

120 plots rendered linear trends for all different reactants. The result-
ing adsorption and rate constants obtained by this method are
100 summarized in Table. 3.
m-Cr To further validate this method, the kinetics and adsorption
ONE parameters obtained in the differential reactor for toluene and
80 OL
1/r (g.hr/mol)

methyl-cyclohexane were used to compare the data obtained in


TOL
the integral reactor by using the full LangmuirHinshelwood rate
60 MCH
ENE expressions for toluene and methyl-cyclohexane conversions,
respectively:
40
kTM  K TOL  PTOL
rTM 6
1 K TOL  PTOL K MCH  PMCH
20
kMT  K MCH  PMCH
rMT 7
1 K TOL  PTOL K MCH  PMCH
0
0.00030 0.00045 0.00060 0.00075
(where TM indicates toluene to methyl-cyclohexane and MT
1/P (Pa-1) methyl-cyclohexane to toluene). The adsorption and rate constants
for toluene (KTOL and kTM) as well as for methyl-cyclohexane (KMCH,
Fig. 5. 1/r vs. 1/P plot for different reactants in differential reactor over Pt/SiO2.
kMT) from Table 3 (differential reactor) were directly used with the
Temperature = 300 C. Pressure = 1 atm. Pm-Cr = 1486 Pa. MCH: methyl-cyclohex-
ane; TOL: toluene; m-Cr: m-cresol; ONE: 3-methyl-cyclohexanone; OL: 3-methyl- rate expressions (6) and (7) to model the integral reactor, resulting
cyclohexanol; ENE: 3-methyl-1-cyclohexene. in excellent agreement over the entire W/F range, as shown in Fig. 6.

Table 3 3.4. Goodness of the tting from the integral reactor


Adsorption constants calculated from data in Fig. 5. m-Cr: m-cresol; ONE: 3-methyl-
cyclohexanone; OL: 3-methyl-cyclohexanol; TOL: toluene; MCH: methyl-cyclohex- The adsorption constants obtained from the differential reactor
ane; ENE: 3-methyl-1-cyclohexene.
(KTOL, KMCH as well as KCr, KONE, KOL, and KENE) were used and kept
Reactant Slope Intercept Adsorption constant Rate constant xed in the kinetic tting of the entire reaction pathway
K  1000 (Pa1) k  10 (mol/(g h Pa)) (Scheme 1). Similarly, the rate constants kET and kEM values
m-Cr 32,356 13.6 0.42 obtained from feeding methylcyclohexene (Table 2) were also kept
ONE 863 1.7 1.96 xed. The rest of the rate constants, including kDOH, kDW, kCK, kKC,
OL 2565 1.5 0.58
kKO, kOK, kOC, and kCO, were used as adjustable parameters to
TOL 125,216 30.8 0.25 0.32
MCH 17,416 1.7 0.10 5.8
optimize the tting. Fig. 2(ac) illustrate the agreement between
ENE 24 0.35 0.15 the experimental results of conversion data of m-cresol, 3-
methyl-cyclohexanone (ONE), and 3-methyl-cyclohexanol (OL)
and the simultaneous tting using the same set of parameters.
As mentioned above, the initial values for the adjustable parame-
over the Pt/SiO2 catalyst under differential conditions, at 300 C, for ters used in the tting were chose from those obtained from the
three different partial pressures. As shown in Fig. 5, the 1/r vs. 1/P low conversion region in Fig. 2. The nal optimized values obtained
28 L. Nie, D.E. Resasco / Journal of Catalysis 317 (2014) 2229

(a) 100 (b) 100

Product Distribution (mol %)

Product Distribution (mol %)


90
R2 = 0.98
80 80
R2 = 0.99
70
60 MCH 60 MCH
TOL TOL
50 MCH Fitting
MCH Fitting
40 TOL Fitting 40 TOL Fitting
30
20 20
10
0 0
0 2 4 6 8 10 12 14 16 0 2 4 6 8 10 12 14
W/F (gcat.hr/mol) W/F (gcat.hr/mol)

Fig. 6. Product distribution from reaction of methyl-cyclohexane and toluene over Pt/SiO2. (a) methyl-cyclohexane, (b) toluene. Temperature = 300 C. Pressure = 1 atm.
PMCH = PTOL = 1486 Pa. MCH: methyl-cyclohexane; TOL: toluene. The points are experimental data and the lines are calculated from the kinetics model (Eqs. (6) and (7)), using
the individual thermodynamic and kinetics constants obtained in the differential reactor for each specic feed (MCH or TOL).

were found to be within 20% of the initial inputs. As shown in lower rate constant kDW, is via hydrogenation of m-cresol to the
Fig. 2, the solid lines calculated from kinetic model t the experi- corresponding alcohol followed by dehydration.
mental data points very well. The rate constants resulting from Another important demonstration that the hydrogenation/
the tting are listed in Table 4. dehydration route occurs only to a minor extent on Pt/SiO2 is
The small rate constant obtained in the tting for the 3-methyl- obtained from the analysis of the evolution of the toluene/methyl-
cyclohexanol dehydration (kDW = 0.004 mol/(g h)) corresponds well cyclohexane product ratio. Table 2 shows the results of the conver-
with that directly measured over pure SiO2 under the same reaction sion of 3-methyl-1-cyclohexene to toluene and m-cyclohexane
condition, (kDW = 0.005 mol/(g h)) as shown in Table 2. This good over Pt/SiO2 at 300 C. In the low conversion range (12.6% and
agreement gives further support to the conclusion that the contri- 18.9%), the reaction is under kinetic control and the observed
bution of the HDO route to the deoxygenation of m-cresol on Pt/ toluene/methylcyclohexane ratios are 0.87 and 1.2, respectively.
SiO2 is very small. Also, as previously mentioned in a more qualita- Therefore, if during m-cresol, deoxygenation toluene and methyl-
tive way, the interconversion of m-cresol, 3-methyl-cyclohexanone cyclohexane would primarily arise from dehydration of 3-
(ONE), and 3-methyl-cyclohexanol (OL) is signicantly faster than methyl-cyclohexanol to 3-methyl-1-cyclohexene, the subsequent
hydrodeoxygenation, as quantied by the relative values of the rate dehydrogenation/hydrogenation would produce toluene and
constants. For instance, kCK (m-Cr ? ONE) and kKC (ONE ? m-Cr) methyl-cyclohexane at a ratio close to that obtained when feeding
are 510 times larger than kDOH (m-Cr ? TOL). 3-methyl-1-cyclohexene alone (i.e., 0.871.2). However, the tolu-
The validity and robustness of the kinetic model and tting ene/methyl-cyclohexane ratio obtained during the conversion of
parameters was further tested by feeding a mixture of m-cresol m-cresol conversion is remarkably higher, for example, about 40
and 3-methyl-cyclohexanol at a 1:1 M ratio over Pt/SiO2. Without at 10% conversion. Moreover, it is suggestive that this value is
changing any of the tting parameters obtained in the original t- far larger than the equilibrium ratio predicted at this temperature
ting, the experimental data were compared with the values pre- and 1 atm of H2, i.e. toluene/methyl-cyclohexane = 11. This is a
dicted from the LH kinetic model. As shown in Fig. 7, a very good clear indication that toluene is formed before methylcyclo-
agreement between the predicted values (solid lines) and the hexane, or even methylcyclohexene. Otherwise, the toluene/
experimental data (symbols) was obtained in both additional runs. methyl-

3.5. Reaction pathway analysis based on the kinetics model


70

The kinetics analysis provides quantitative validation to the R2 = 0.98 MCH


60
reaction pathway shown in Scheme 1. That is, on the Pt/SiO2 cata- TOL
Product distribution (%)

lyst, the dominant deoxygenation path, with a higher rate constant ONE
50
kDOH, takes place via direct dehydroxylation; the other path, with a OL
m-Cr
40 MCH fitting
Table 4
TOL fitting
Rate constant for different reaction steps. m-Cr: m-cresol; ONE: 3-methyl-cyclohex-
30 ONE fitting
anone; OL: 3-methyl-cyclohexanol; TOL: toluene; MCH: methyl-cyclohexane; ENE: 3-
methyl-1-cyclohexene.
OL fitting
20 m-Cr fitting
Rate constant Parameter value (mol/(g h))
kDOH (m-Cr ? TOL) 0.024 10
kDW (OL ? ENE) 0.004
kCK (m-Cr ? ONE) 0.139
0
kKC (ONE ? m-Cr) 0.236 0 1 2 3 4 5
kKO (ONE ? OL) 0.055
kOK (OL ? ONE) 0.441
W/F (gcat.hr/mol)
kOC (OL ? m-Cr) 0.244
kCO (m-Cr ? OL) 0.007 Fig. 7. Product distribution from reaction of mixture of m-cresol and 3-methyl-
kET (ENE ? TOL) 1.70 cyclohexanol (1:1 M ratio) over Pt/SiO2. Temperature = 300 C. Pressure = 1 atm.
kEM (ENE ? MCH) 2.00 MCH: methyl-cyclohexane; TOL: toluene; m-Cr: m-cresol; ONE: 3-methyl-cyclo-
kTM (TOL ? MCH) 0.350 hexanone; OL: 3-methyl-cyclohexanol. The points are experimental data and the
lines are calculated from the LangmuirHinshelwood kinetics model.
L. Nie, D.E. Resasco / Journal of Catalysis 317 (2014) 2229 29

Table 5 tautomerization/secondary hydrogenation route with the HYD


Product selectivity for m-cresol conversion over different support Pt catalysts at route (fully hydrogenation/dehydration/dehydrogenation) only
300 C. Pressure = 1 atm. MCH: methyl-cyclohexane; TOL: toluene; m-Cr: m-cresol;
ONE: 3-methyl-cyclohexanone; OL: 3-methyl-cyclohexanol.
playing a minor role. Moreover, using TiO2 or ZrO2 as the support
instead of SiO2 can greatly improve the hydrodeoxygenation selec-
Catalyst Conversion (%) Selectivity (%) tivity. The enhancement is due to the enhanced chemoselective
TOL ONE OL Benzene MCH hydrogenation of the C@O bond of the reaction intermediate
1%Pt/TiO2 17 88 3 0.0 1 8 3-methyl-3,5-cyclohexadienone (i.e., the m-cresol tautomer).
1%Pt/SiO2 17 28 57 14 0 1
1%Pt/ZrO2 12 67 8 0.4 0.5 21 Acknowledgements

Support from the National Science Foundation (NSF EPS-


cyclohexane ratio could never be higher than the equilibrium
CoR0814361) and the Department of Energy (DE-FG36GO88064)
value.
is greatly appreciated. Allocation of computing time provided by
the OU Supercomputing Center for Education and Research
3.6. Variation of HDO selectivity by effect of the support
(OSCER) at the University of Oklahoma is acknowledged. We thank
Prof. Tawan Sooknoi (King Mongkuts Institute of Technology
We have observed here that the selectivity of Pt catalyst to each
Ladkrabang, Thailand) for valuable discussions.
of the possible paths can be controlled by the choice of support. As
shown in Table 5, the product selectivities (measured at similar
References
conversion levels) over different supported Pt catalysts greatly
depend on the type of support used. That is, when the support is [1] Q. Bu, H. Lei, A.H. Zacher, L. Wang, S. Ren, J. Liang, Y. Wei, Y. Liu, J. Tang, Q.
TiO2 or ZrO2, the selectivity to toluene was greatly enhanced, com- Zhang, R. Ruan, Bioresour. Technol. 124 (2012) 470477.
pared to Pt/SiO2, over which 3-methyl-cyclohexanone (ONE) was [2] J.P. Diebold, A review of the chemical and physical mechanisms of the storage
stability of fast pyrolysis bio-oils (2000). (NREL/SR-570-27613).
dominant. For the former, the main secondary product was meth- [3] D.M. Alonso, J.Q. Bond, J.A. Dumesic, Green Chem. 12 (2010) 14931513.
ylcyclohexane, while for the latter, it was 3-methyl-cyclohexanol [4] T. Pham, D. Shi, D.E. Resasco, Appl. Catal. B 145 (2014) 1023.
(OL). This difference can be ascribed to the chemoselective [5] A. Ausavasukhi, Y. Huang, A.T. To, T. Sooknoi, D.E. Resasco, J. Catal. 290 (2012)
90100.
hydrogenation of the reaction intermediate 3-methyl-3,5-cyclo- [6] S. Boonyasuwat, T. Omotoso, D.E. Resasco, Catal. Lett. 143 (8) (2013) 783791.
hexadienone (i.e., the m-cresol tautomer). As mentioned above [7] D.C. Elliott, Energy Fuels 21 (2007) 17921815.
when the C@C double bonds of the keto-isomer are hydrogenated, [8] D.C. Elliott, T.R. Hart, G.G. Neuenschwander, L.J. Rotness, M.V. Olarte, A.H.
Zacher, Y. Solantausta, Energy Fuels 26 (2012) 38913896.
the primary product is 3-methyl-cyclohexanone. By contrast, when [9] D.C. Elliott, WIREs, Energy Environ. 2 (2013) 525533, http://dx.doi.org/
the carbonyl group of the ketone intermediate is hydrogenated, the 10.1002/wene.74.
product is 3-methyl-3,5-cyclohexadienone, which can be easily [10] M.J. Girgis, B.C. Gates, Ind. Eng. Chem. Res. 30 (1991) 20212058.
[11] B. Feng, H. Kobayashi, H. Ohta, A. Fukuoka, J. Mol. Catal. A: Chem (2013).
dehydrated to toluene as a deoxygenation product. As previously
http://dx.doi.org/10.1016/j.molcata.2013.09.025.
reported by Vannice et al. [38] for the hydrogenation of crotonalde- [12] X. Zhu, L.L. Lobban, R.G. Mallinson, D.E. Resasco, J. Catal. 281 (2011) 2129.
hyde, Pt/SiO2 has a high preference for hydrogenating C@C double [13] M. Saidi, F. Samimi, D. Karimipourfard, T. Nimmanwudipong, B.C. Gates, M.R.
Rahimpour, Energy Environ. Sci. 7 (2014) 103.
bonds. However, Pt/TiO2 results in higher selectivity for C@O
[14] D.C. Elliott, T.R. Hart, Energy Fuels 23 (2009) 631637.
hydrogenation, which is due to the strong interaction and stabil- [15] C. Zhao, Y. Yu, A. Jentys, J.A. Lercher, Appl. Catal. B 132133 (2013) 282292.
ization of carbonyl group at interfacial PtTiOx sites which are oxo- [16] C. Zhao, S. Kasakov, J. He, J.A. Lercher, J. Catal. 296 (2012) 1223.
philic. Thus, over Pt/TiO2 as well as over Pt/ZrO2, for a given [17] T. Nimmanwudipong, R.C. Runnebaum, D.E. Block, B.C. Gates, Energy Fuels 25
(8) (2011) 34173427.
conversion level, the toluene yield is much higher than on Pt/ [18] H. Wang, J. Male, Y. Wang, ACS Catal. 3 (2013) 10471070.
SiO2. The dramatic difference in product distribution for different [19] F.E. Massoth, J. Simons, J. Phys. Chem. B 110 (2006) 1428314291.
catalyst supports further proves that the hydrodeoxygenation of [20] A.J. Foster, P.T.M. Do, R.F. Lobo, Top Catal. 55 (2012) 118128.
[21] R. Prins, Handbook of Heterogenous Catalysis, pp. 26962718.
m-cresol under the reaction conditions in this contribution is via [22] E.O. Odebunmi, D.F. Ollis, J. Catal. 80 (1983) 5664.
the tautomerization/secondary chemoselective hydrogenation [23] E. Furimsky, Apply Catal. A: Gen. 199 (2000) 147190.
pathway. [24] B. Gvenatama, O. Kursun, E.H.J. Heeresb, E.A. Pidko, E.J.M. Hensen, Catal.
Today (2013). http://dx.doi.org/10.1016/j.cattod.2013.12.011.
[25] J. He, C. Zhao, J.A. Lercher, J. Catal. 309 (2014) 362375.
4. Conclusions [26] B. Feng, H. Kobayashi, H. Ohta, A. Fukuoka, J. Mol. Catal. A: Chem. (2013).
http://dx.doi.org/10.1016/j.molcata.2013.09.025.
[27] P.M. Mortensen, J.-D. Grunwaldt, P.A. Jensen, A.D. Jensen, ACS Catal. 3 (2013)
Hydrodeoxygenation (HDO) of m-cresol, a model phenolic 17741785.
compound for fast pyrolysis bio-oil, over Pt/SiO2 produces toluene [28] B. Yoosuk, D. Tumnantong, P. Prasassarakich, Chem. Eng. Sci. 79 (2012) 17.
as the major deoxygenation product, while the conversion of 3- [29] M. Badawi, J.-F. Paul, E. Payen, Y. Romero, F. Richard, S. Brune, A. Popov, E.
Kondratieva, J.-P. Gilson, L. Mariey, A. Travert, F. Maug, Oil Gas Sci Technol. 5
methyl-cyclohexanone and 3-methyl-cyclohexanol, respectively, (2013) 829840.
over the catalyst only produces minor toluene and methyl- [30] B.S. Gevert, J.-E. Otterstedt, F.E. Massoth, Appl. Catal. 31 (1987) 119131.
cyclohexane. Fine LangmuirHinshelwood kinetic tting with [31] L. Nie, P.M. de Souza, F.B. Noronha, W. An, T. Sooknoi, D.E. Resasco, J. Mol.
Catal. A: Chem. 388389 (2014) 4755.
numerous parameters could be obtained by delicately combining [32] R.J. Madon, M. Boudart, Ind. Eng. Chem. Fund. 21 (1982) 438.
a differential reactor and an integral reactor. The obtained kinetic [33] E.O. Odebunmi, D.F. Ollis, J. Catal. 80 (1983) 7689.
parameters (rate constants) show that the hydrogenation/dehy- [34] S. Sitthisa, T. Sooknoi, Y. Ma, P.B. Balbuena, D.E. Resasco, J. Catal. 277 (2011) 1
13.
drogenation between the reaction intermediates takes place at a [35] R.M. Watwe, R.D. Cortright, J.K. Norskov, J.A. Dumesic, J. Phys. Chem. B 104
faster rate than deoxygenation. The adsorption of both the reac- (2000) 22992310.
tants and the products depresses the reaction rates. Here, over [36] R.D. Cortright, J.A. Dumesic, Adv. Catal. 46 (2001) 161264.
[37] M.A. Vannice, in: Kinetics of Catalytic Reactions, Springer, Science, N. York,
nonacidic Pt/SiO2, under the reaction condition, the deoxygenation
2005. ISBN:1441937587.
of m-cresol does not require a pre-hydrogenation of the aromatic [38] M.A. Vannice, B. Sen, J. Catal. 115 (1989) 6578.
ring. It is demonstrated that the reaction mainly occurs via the

Das könnte Ihnen auch gefallen