Vous êtes sur la page 1sur 5

Journal of Molecular Catalysis A: Chemical 368369 (2013) 6165

Contents lists available at SciVerse ScienceDirect


Journal of Molecular Catalysis A: Chemical
j our nal home page: www. el sevi er . com/ l ocat e/ mol cat a
Green and efcient synthesis route of catechol from guaiacol
Le Yang
a
, Wei Zhou
a
, K. Seshan
b,1
, Yongdan Li
a,
a
Tianjin Key Laboratory of Applied Catalysis Science and Technology and State Key Laboratory of Chemical Engineering (Tianjin University), School of Chemical Engineering, Tianjin
University, Tianjin 300072, China
b
Faculty of Science & Technology, Catalytic Process and Materials, University of Twente, P.O. Box 217, 7500 AE, Enschede, The Netherlands
a r t i c l e i n f o
Article history:
Received 27 August 2012
Received in revised form
11 November 2012
Accepted 28 November 2012
Available online 7 December 2012
Keywords:
Catechol
HCl
High-temperature water
Guaiacol
Lignin
a b s t r a c t
The conversion of guaiacol to catechol in high temperature water by catalysis of mineral hydrochloric
acid is examined. The effects of pHand H
2
pressure are measured in the conversion of guaiacol. Hydrogen
enhances the reaction dramatically. Moreover, low pH and high hydrogen pressure favor the reaction.
The highest conversion of guaiacol and best yield of catechol of 99% and 89%, respectively, are achieved
with 1 MPa hydrogen at pH = 1.8 for 3h at 280

C. Based on the experimental results and kinetics, possible


reaction mechanisms are proposed. Besides ionic mechanism and water catalysis, hydrogen polarization
also occurred without metal catalyst.
2012 Elsevier B.V. All rights reserved.
1. Introduction
Biomass has drawn an increasing attention in recent years
as a renewable feedstock in response to the depletion of fossil
fuels and the increasing concern over environmental protection.
Woody biomass consists of 4045wt.% cellulose, 2535wt.% hemi-
cellulose, 1530wt.% lignin, and other minor components [1].
Signicant achievements have already been made in the produc-
tion of fuels and chemicals, such as sugars, alcohols and furans,
from cellulose and hemicellulose, leaving lignin still as a waste
[27]. In addition, pulp and paper industries also generate a huge
amount of waste lignin which causes pollution in quite some cases.
Presently, for the better case, lignin is used as a low-grade boiler
fuel to provide heat and power. Less than 5% of the worlds sup-
ply of lignin has been used for other purposes. Between 40 and
50 million tons per annum are produced worldwide as a mostly
non-commercialized waste product.
Lignin is a highly branched polymer of substituted phenolic
monomeric entities. Reactions of lignin are very complex and of
low yield, due to its amorphous structure and easy condensibility
of its monomers, leaving lignin underutilized to its potential [8].
Guaiacol, a typical monomeric species found in lignin, has been
studied widely for the understanding of lignin chemistry [916].

Corresponding author. Fax: +86 22 27405243.


E-mail address: ydli@tju.edu.cn (Y. Li).
1
Fax: +31 53 4893254/4893033.
Guaiacol can be converted to catechol in organic solvents [913]
and water [1] especially high temperature water (HTW) and super-
critical water (SCW).
Generally, there are two kinds of products for guaiacol conver-
sion, i.e. chemical and fuel. Homogenous catalysts such as FeCl
3
and RuCl
3
[9], heterogeneous catalysts such as Al
2
O
3
[10], Re/ZrO
2
[11], Mo
2
N [12], and transition metal phosphide [13] were applied
to get chemicals such as phenol, cresol, benzene, catechol and
ethoxyphenol. However, the conversion of guaiacol and the yield
of chemicals are moderate. Catalytic hydrodeoxygenation (HDO)
of guaiacol at high pressure and temperature is used to achieve
fuel. Sulde catalysts such as sulded CoMo and NiMo [15], pre-
cious metal catalysts such as Pd, Ru [14] and bimetallic catalysts
[16] have been explored. However, it is less competitive to con-
vert lignin to fuels than to chemicals because the H/C ratio in lignin
is rather low and lignin contains oxygen. According to Bridgwater
[17], 62kg of H
2
is required to hydrodeoxygenate one ton of wood-
based pyrolysis oil to fuel grade material. Nevertheless, people can
make great advantage of lignin to produce phenolic compounds
rather than fuels.
Catechol is one of such products of much value as chemical or
chemical intermediate, which is an important chemical that can
be used directly or used as a precursor. Approximately 50% of syn-
thetic catechol is used for the production of pesticides, the rest
being used as a precursor to ne chemicals such as perfumes and
pharmaceuticals [18,19]. For example, catechol, as a raw mate-
rial, is applied in the production of avors and fragrances. The
related monoethyl ether of catechol, ethoxyphenol, is converted to
1381-1169/$ see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.molcata.2012.11.024
62 L. Yang et al. / Journal of Molecular Catalysis A: Chemical 368369 (2013) 6165
ethylvanillin which has been used for the production of chocolate
confectioneries [16,20]. Catechol is also used as a black-and-white
photographic developer.
A few synthetic procedures exist for transformation of sub-
stituted phenols into catechols. One practical procedure involves
ortho-formylation of phenols followed by a subsequent Dakin
oxidation [21]. However, these processes suffer from low selec-
tivity particularly for meta-substituted phenols. Another method
employs oxidation of phenol to o-quinone and a subsequent reduc-
tionof the latter into catechol. The industrial versionof this method
employing H
2
O
2
oxidation, in which catechol is produced by the
hydroxylation of phenol, usually provides a mixture of catechol
and para-hydroquinone [22]. An efcient, green, sustainable and
selective route for catechol is demanded.
In addition, production of high-value chemicals can be the eco-
nomic driver of lignin research [23].
Therefore, we propose a green and sustainable route
ligninguaiacolcatechol of lignin utilization. Guaiacol always
holds a high selectivity among the products of lignin depolymer-
ization [24,25], so we chose guaiacol as the precursor of catechol.
In this work, we focus on the second half route guaiacolcatechol
and aim to produce catechol from guaiacol with a high yield. Ef-
cient conversion of lignin-based guaiacol to catechol is reported
here. In this context, close to 90% catechol yields, obtained for the
rst time is reported in this study.
2. Experimental
2.1. Materials
Guaiacol, catechol, phenol, cresol, ethylbenzene, ethanol, con-
centrated hydrochloric acid, FeCl
3
6H
2
O and CuCl
2
2H
2
O were
obtained fromGuangfu Inc, all with purity higher than 99%. Guaia-
col was used as a starting material. Concentrated hydrochloric acid,
FeCl
3
6H
2
O and CuCl
2
2H
2
O were used as catalysts. Catechol, phe-
nol, cresol were used to calibrate the GC. Ethylbenzene was chosen
as an internal standard of the GC analysis, and ethanol was used to
help the dissolving of ethylbenzene.
2.2. Reaction condition
The experiments were carried out in a batch reactor (Parr
4566, made of Hastelloy), with an internal volume 300cm
3
and
with a temperature controller (Parr 4848). In the reactions with
hydrochloric acid, a pH meter calibrated at pH 4.0 with a com-
mercial buffer solution was used. First, hydrochloric acid solution
with desired pH was made in a beaker. The pH values reported
for the reaction experiments in this article are those calculated to
exist at reaction conditions. The pH of neutral water was 5.6 at
280

C, and the H
+
concentrations were calculated based on the
complete dissociation of the acid [26]. The reactor was loaded with
70mL hydrochloric acidic solution and 0.5g guaiacol. After sealing
the reactor, nitrogen was used to purge the reactor, followed by
hydrogen if necessary. Then the reactor was heated to the desired
temperature 280

C, and the reaction was carried out for 3h. The


initial gas phase composition, H
2
or N
2
, were examined. The ini-
tial pressures reported in this work are the gauge values measured
at ambient temperature and the overall reaction pressures are
increased with the increase of the temperature.
In the reactions with FeCl
3
and CuCl
2
, rst 70mL deionized
water was added into a beaker and then a catalyst (mole ratio
between catalyst and guaiacol was 0.45). After the complete dis-
solution by ultrasonic treatment, the pH of the solution was
measured. Then the solution and 0.5g guaiacol were added into the
reactor. The next steps were the same as the reactions catalyzed by
hydrochloric acid. The initial atmosphere is 0MPa H
2
.
2.3. Analysis
Ethylbenzene was chosen as an internal standard and
was added into the product, ethanol was also added to
help the dissolving of ethylbenzene. A GCMS (Agilent 6890-
5973, 30m0.25mm0.25m HP-5MS capillary column) was
used for products identication and a GCFID (Agilent 6890,
50m0.2mm0.25m HP-5 capillary column) was used for
quantitative analysis.
The conversion of guaiacol and the yield of catechol were
dened as follows: conversion=[(mole of guaiacol)
in
(mole
of guaiacol)
out
]/(mole of guaiacol)
in
100%, yield=mole of cat-
echol)/(mole of guaiacol)
in
100%. The pseudo-rst-order rate
constant for guaiacol hydrolysis was calculated as: k =ln(1x)/t
3. Results and discussion
3.1. Effect of pH and metal chlorides
The effect of pH and the addition of chlorides on the conver-
sion and yield are shown in Fig. 1. Both guaiacol conversion and
Fig. 1. Effect of pH (manipulated with addition of HCl), addition of metal chloride with water as the solvent and the gas atmosphere on guaiacol conversion and catechol
yield with an initial pressure (gauge) 0MPa at 280

C for 3h.
L. Yang et al. / Journal of Molecular Catalysis A: Chemical 368369 (2013) 6165 63
catechol yield increase at lower pH values. Moreover, the presence
of hydrogen enhances the guaiacol conversion and catechol yields
dramatically. With the presence of hydrogen, the highest conver-
sion and yield are 83% and 76% at pH 1.8. On the other hand, it is
obvious from Fig. 1 that both the conversion and yield are much
higher in hydrogen than in nitrogen atmosphere. The selectivity
of catechol in these experiments is as high as 90% irrespective to
the gas phase composition. The results with metal chlorides (FeCl
3
,
CuCl
2
) as catalysts in hydrogen atmosphere were also shown in
Fig. 1. The metal chlorides induce an acidic environment which
favors the reaction. The pH of FeCl
3
and CuCl
2
solutions were 2.6
and3.8, respectively. The yieldwithFeCl
3
is higher thanwithCuCl
2
.
However, the yields in the reactions with the two metal chlorides
are substantially higher than the cases with only HCl at the similar
pH values, which indicates that the metal sites are involved in the
reaction.
3.2. Effect of pH on the reaction kinetics
The rate constants were calculated based on a rst-order kinet-
ics (Fig. 2) for the acid catalyzed reaction. If the reaction is truly
catalyzed by H
+
, the pseudo-rst-order reaction rate constant, k,
for the disappearance of guaiacol should be directly proportional to
the ion concentration in the solution. More specically the slope of
logk vs. pHshouldbe 1as wouldtypical for H
+
catalyzedreactions
[26].
In this work, the slope of the solid line drawn through the data
between pH 1.8 and 3.9 is about 1, which means the apparent
reaction order for H
+
is 1.0, same as that would be expected for a
reaction following specic acid catalysis. Nevertheless, as the pH
continues to increase from 3.9 to 5.8, the slope of the dash dot
line is about 0.25 instead of 1.0, indicating the reaction in this pH
range is not specically acid catalysis. Savage et al. proposed that
catalysis by water is a second possible pathway to generate the
1 2 3 4 5 6
-6.5
-6.0
-5.5
-5.0
-4.5
-4.0
-3.5
H2
N2
l
o
g
(
k
/
s
-
1
)
pH
Fig. 2. Effect of pH on the hydrolysis reaction rate constants at 280

C.
initial protonated intermediates in the hydrolysis of dibenzyl ether
[26]. It is known that the ion product increases by 34 orders of
magnitude from its value at ambient conditions to its maximum
value at around 250

C. Simply increase the temperature of the


liquid water can profoundly increase the ion concentrations. This
enhanced acidity/basicity from the increased ion product has led
researchers to explore HTWas a mediumfor acid-/base-catalyzed
reactions. Besides, Hunter et al. also demonstrated that a conven-
tional general acid catalysis mechanism cannot account for the
inuence of pHon the reaction kinetics of tetrahydrofuran synthe-
sis in HTW[27]. In accordance with these suggestions we propose
that water molecules, rather than H
+
, play the role of a catalyst
between pH 3.9 and 5.8.
The mechanism of guaiacol hydrolysis is put forward in
Scheme 1 according to the experimental facts and the inspiration
by the work reported in reference 26. Guaiacol is cleaved by S
N
2
Scheme 1. Guaiacol hydrolysis pathways ionic mechanism.
64 L. Yang et al. / Journal of Molecular Catalysis A: Chemical 368369 (2013) 6165
0.0 0.2 0.4 0.6 0.8 1.0
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
yield at pH 5.8
yield at pH 2.0
yield at pH 1.8
conversion at pH 5.8
conversion at pH 2.0
conversion at pH 1.8
pressure/MPa
Fig. 3. Effect of initial hydrogen pressure on the conversion of guaiacol and the yield
of catechol at pH 1.8, 2.0 and 5.8 with initial hydrogen pressure of 0MPa, 0.5MPa,
1 MPa at 280

C for 3h.
mechanismand an ionic pathway dominates the reaction. Step 1a
demonstrates an acid-catalyzed mechanismat a lowpH, while step
1b demonstrates water-catalyzed mechanismat a higher pH. Pro-
tonation of guaiacol in both step 1a and step 1b forms a charged
intermediate, which subsequently reacts with a water molecule to
eventually formcatechol, methanol and H
+
in step 2.
Step 1a is a process in which a smaller charge-to-volume ratio
intermediate is formed. Savage suggested that with the decrease
of solvent dielectric constant the formation of ions with smaller
charge-to-volume ratio is increasingly favored [28]. Water dielec-
tric constant decreases as the temperature increases. For example,
the dielectric constant is 21 at 300

C and 4.1 at 500

C. Step 2 is
irreversible and rate-determining. H
+
released fromstep 2 goes to
two separate ways, of which one returns to step 1a to act as acid
catalyst and the other neutralizes with the OH

released in step 1b
to formwater acting as the catalyst for step 1b.
3.3. Effect of hydrogen pressure
The inuence of hydrogen pressure on the reaction is show in
Fig. 3. It depicts the effect of H
2
pressure on the conversion and
yield at pH 1.8, 2.0 and 5.8. It is obvious that hydrogen effec-
tively enhances the reaction. The highest conversionreaches 99%at
pH=1.8 with a yield of catechol as high as 90%. This result is much
higher compared to other guaiacol conversion reactions [912]
reported earlier in which the highest yield of catechol is no more
than 50%. Here we almost get nearly complete yield.
Thus, it is beyond doubt that hydrogen indeed takes part in the
reaction based on the fact that not only higher hydrogen pressure
benets the reaction (Fig. 2) but also hydrogen exceeds nitrogen on
motivating the reactionwithfaster reactionrate (Fig. 3). Inaddition
to Scheme 1, we propose here a reaction pathway (Scheme 2) to
illustrate howhydrogentakes part inthe reactionbased ona newer
route(shownas inScheme3) intheconversionsequences proposed
by Walling and Bollyky [29]. Scheme 3 presents that hydrogen is
polarized by alkyl cations. In our reaction, the polarity of reaction
H
O
OH
OH
OH
+ CH
4
+ H
+
+ H
2
Scheme 2. Polarization of H
2
by the reaction intermediate.
R
+
+ H
2
RH + H
+
Scheme 3. Polarization of H
2
by R
+
.
Fig. 4. Gas mass spectrometry of gas products.
intermediate which is shown in Scheme 1 in the solution is much
stronger than alkyl cations, so it makes sense that hydrogen takes
part in the reaction with polarization by the intermediate. In the
other word, the intermediate not only reacts with water to form
catechol and methanol shown in Scheme 1 but also reacts with
hydrogen to form catechol and methane shown in Scheme 2. In
addition, we identied gas products with a gas mass spectrometer
(HIDEN HPR20) and nd CH
4
is indeed produced (Fig. 4).
3.4. The relative selectivity of phenol and catechol
Fig. 5 depicts the effect of atmosphere and pH on the relative
selectivity of phenol and catechol. As the pH increases from 1.8
to neutral 5.8, the phenol/catechol mole ratio steadily increases
in the product. The formation of phenol can be attributed to a
radical reaction mechanism that is guaiacol splitting into phenol
and methanol. Methanol has been identied by GCMS. On the
other hand, compared to hydrogen, nitrogen benets phenol for-
mation dramatically. In the other words, nitrogen favors radical
reaction while hydrogen favors ionic reaction, which can further
verify Scheme 2 that hydrogen participates in ionic reaction.
1 2 3 4 5 6
0.00
0.05
0.10
0.15
0.20
0.25
0.30

p
h
e
n
o
l
/
c
a
t
e
c
h
o
l

m
o
l
e

r
a
t
i
o
in N
2
in H
2
pH
Fig. 5. Effect of atmosphere and pHon the relative selectivity of phenol to catechol.
L. Yang et al. / Journal of Molecular Catalysis A: Chemical 368369 (2013) 6165 65
3.5. Effect of NaOH on the reaction
Hydrolysis of guaiacol in NaOH solution at pH=10.4 (measured
at ambient condition) was tested, and the other reaction condi-
tions were the same as acidic reactions (280

C, 3h, N
2
). The results
showed that basic solution did catalyze the reaction of the pro-
duction of catechol but produces a large amount of undesirable
char visually, in contrast to almost no char formation with acid
catalysis. This phenomenon is in accordance with the hydrolysis of
polysaccharides forming simple sugars such as glucose, fructose,
and xylose, to which acid hydrolysis is more commonly practiced
because base hydrolysis leads tomore side reactions andthus lower
yields [23].
4. Conclusions and perspective
This work shows an efcient and novel chemical sequence to
produce green catechol from guaiacol, an intermediate of bio-
waste, lignin. The route ligninguaiacolcatechol is feasible. High
conversion of guaiacol and yield of catechol were obtained at 99%
and 89%, respectively with 1MPa hydrogen at pH=1.8 for 3h at
280

C. Compared to nitrogen, hydrogen motivated the reaction


dramatically. As a whole, higher hydrogen pressure and lower pH
favor the reaction. Based on the experimental results and kinetics,
possible reaction mechanisms are proposed.
The next task is to maximize guaiacol from lignin depolymer-
ization, it is already the highest phenolic component in the product
obtained by the thermal chemical process of biomass [17,18]. How-
ever, presently the guaiacol yield from lignin depolymerization is
really low and the total phenols are no more than 10% [25,30].
The best overall yield of mixed phenols, 2535%, is known to be
achieved by Kleinert and Barth, who carried out experiments with
lignin in the mixture of formic acid, ethanol and 2-propanol at
380

C for 16h [31].


Acknowledgment
The nancial support fromthe Ministry of Science and Technol-
ogy of China under contract number 2011DFA41000 is gratefully
acknowledged.
References
[1] T. Wahyudiono, M. Kanetake, M Goto Sasaki, Chem. Eng. Technol. 30 (2007)
11131122.
[2] X.L. Tong, M.R. Li, N. Yan, Y.M.P.J. Dyson, Y.D. Li, Catal. Today 175 (2011)
524527.
[3] X.L. Tong, Y. Ma, Y.D. Li, Appl. Catal. A 385 (2010) 113.
[4] X.L. Tong, Y.D. Li, ChemSusChem3 (2010) 350355.
[5] G.W. Huber, J.W. Shabaker, J.A. Dumesic, Science 300 (2003) 20752077.
[6] R.D. Cortright, R.R. Davada, J.A. Dumesic, Nature 418 (2002) 964967.
[7] G.W. Huber, J.N. Chheda, C.J. Barrett, J.A. Dumesic, Science 308 (2005)
14461450.
[8] M.P. Pandey, C.S. Kim, Chem. Eng. Technol. 34 (2011) 2941.
[9] A.I. A, E. Chornet, R.W. Thring, R.P. Overend, Fuel 75 (1996) 509516.
[10] L. Yang, K. Seshan, Y.D. Li, Catal. Commun. 30 (2013) 3639.
[11] C. Sepulveda, K. Leiva, R. Garcia, L.R. Radovic, I.T. Ghampson, W.J. Desisto, J.L.
Garcia Fierro, N. Escalona, Catal. Today 172 (2011) 232239.
[12] P.E. Ruiz, K. Leiva, R. Garcia, P. Reyes, J.L.G. Fierro, Appl. Catal. A 384 (2010)
7883.
[13] H.Y. Zhao, D. Li, P. Bui, S.T. Oyama, Appl. Catal. A 391 (2011) 305310.
[14] D.C. Elliott, T.R. Hart, Energy Fuels 23 (2009) 631637.
[15] P.E. Ruiz, B.G. Frederick, W.J. De Sisto, R.N. Austin, L.R. Radovic, K. Leiva, R.
Garca, N. Escalona, Catal. Commun. 27 (2012) 4448.
[16] C.R. Lee, J.S. Yoon, Y.W. Suh, J.W. Choi, J.M. Ha, D.J.Y.K. Suh, Park Catal. Commun.
17 (2012) 5458.
[17] A.V. Bridgwater, Catal. Today 29 (1996) 285295.
[18] H. Fiegel, H.W. Voges, T. Hamamoto, S. Umemura, T. Iwata, H. Miki, Y. Fujita, H.J.
Buysch, D. Garbe, W. Paulus, Ullmanns Encyclopedia of Industrial Chemistry,
Wiley-VCH, Weinheim, 2002.
[19] B.A. Barner, Encyclopedia of Reagents for Organic Synthesis, J. Wiley & Sons,
2004.
[20] K.G. Fahlbusch, F.J. Hammerschmidt, J. Panten, W. Pickenhagen, D. Schatkowski,
K. Bauer, D. Garbe, H. Surburg, Ullmanns Encyclopedia of Industrial Chemistry,
Wiley-VCH, Weinheim, 2005.
[21] T.V. Hansen, L. Skattebol, Tetrahedron Lett. 46 (2005) 33573358.
[22] E.A. Karakhanov, A.L. Maximov, Y.S. Kardasheva, V.A. Skorkin, S.V. Kardashev,
E.A. Ivanova, E. Lurie-Luke, J.A. Seeley, S.L. Cron, Ind. Eng. Chem. Res. 49 (2010)
46074613.
[23] J.N. Chheda, G.W. Huber, J.A. Dumesic, Angew. Chem. Int. Ed. 46 (2007)
71647183.
[24] S. Karagoz, T. Bhaskar, A. Muto, Y. Sakata, Fuel 84 (2005) 875884.
[25] Z.S. Yuan, S.N. Cheng, M. Leitch, C.B. Xu, Bioresour. Technol. 101 (2010)
93089313.
[26] C.M. Comisar, S.E. Hunter, A. Walton, P.E. Savage, Ind. Eng. Chem. Res. 47 (2008)
577584.
[27] S.E. Hunter, C.E. Ehrenberger, P.E. Savage, J. Org. Chem. 71 (2006) 2296239.
[28] N. Akiya, P.E. Savage, Chem. Rev. 102 (2002) 27252750.
[29] C. Walling, L. Bollyky, J. Am. Chem. Soc. 83 (1961) 29682969.
[30] R.A. Gosselink, W. Teunissen, J.G. van Dam, E.D. Jong, G. Gellerstedt, E.L. Scott,
J.M. Sanders, Bioresour. Technol. 106 (2012) 173177.
[31] M. Kleinert, T. Barth, Chem. Eng. Technol. 31 (2008) 736745.

Vous aimerez peut-être aussi