You are on page 1of 8

Green Chemistry

View Article Online


PAPER View Journal | View Issue
Published on 19 June 2014. Downloaded by Federal University of Minas Gerais on 24/04/2015 15:03:26.

High-yield production of levulinic acid from


Cite this: Green Chem., 2014, 16,
cellulose and its upgrading to -valerolactone
3846
Daqian Ding, Jianjian Wang, Jinxu Xi, Xiaohui Liu, Guanzhong Lu* and Yanqin Wang*

Direct catalytic conversion of cellulose to levulinic acid (LA) by niobium-based solid acids and further
upgrading to -valerolactone (GVL) on a Ru/C catalyst were realized through sequential reactions in a
reactor. Firstly, using aluminium-modied mesoporous niobium phosphate as a catalyst, cellulose can be
directly converted to LA with as high as 52.9% yield in aqueous solution, even in the presence of the Ru/C
catalyst. To the best of our knowledge, this is the best result over a heterogeneous catalyst so far. It was
found that the type of acid (Lewis and Brnsted acids) and acid strength had an inuence on the yield of
LA; the doping of aluminium can enhance the strong Lewis and Brnsted acids, especially the strong
Lewis acid, thus resulting in the increase of LA yield from cellulose as well as from glucose and HMF. Such
an enhancement by a Lewis acid on LA yield from HMF was further conrmed by adding lanthanum
triuoroacetate [(TfO)3La], a strong Lewis acid, in the catalytic system (HCl, (TfO)3H, niobium phosphate),
Received 24th April 2014, indicating that a suitable ratio of Lewis/Brnsted acid is important for higher selectivity to LA from HMF, as
Accepted 1st June 2014
well as from cellulose. Then, after replacing N2 with H2, the generated LA in the reaction mixture can
DOI: 10.1039/c4gc00737a be directly converted to -valerolactone through hydrogenation over the Ru/C catalyst without further
www.rsc.org/greenchem separation of LA.

Introduction Generally, LA is produced from biomass under acidic con-


ditions or with acid catalysts via two approaches: furfuralcohol
With the deterioration of the environment and the inevitable hydrolysis10 and direct biomass conversion.11 In the second
depletion of fossil resources, renewable and abundant approach, the feedstocks could be glucose, starch or cellulose.
biomass is considered to be a promising alternative to non- As the cheapest and most abundant feedstock, cellulose, com-
renewable natural resources for sustainable biofuels and bio- posed of glucose units linked by -1,4-glucosidic bonds, was
chemicals in the future.13 Recently, extensive research studies firstly converted into water-soluble glucose with acidic
have focused on the conversion of biomass into fuels and catalysts, then the generated glucose was dehydrated into
chemicals. Among these explorations, an approach for convert- 5-hydroxymethylfurfural (HMF), and subsequently the formed
ing carbohydrate to levulinic acid (LA) is very important, HMF was further converted into LA and formic acid through
because as one of the 12 high-potential and versatile chemical the hydration and rearrangement process under the same
platform compounds,4 LA, as well as its derivatives, is widely acidic conditions (Scheme 1).
used in medicine, agriculture, food, cosmetics, and spice The chemical degradation of cellulose has been investi-
industries, and biofuels probably in future. For example, levu- gated extensively,12 and many of the studies were conducted
linate esters can be used as solvents in industrial extraction, in strongly acidic aqueous solutions1317 or in the presence of
spices in cosmetics and medicine for injection.5,6 Moreover, ionic liquids,1820 due to the robust structure of cellulose.
-valerolactone (GVL), produced through selective hydrogen- Therefore, preparation of the solid acid catalyst with higher
ation of LA or levulinate esters, is also considered as an impor- activities under mild conditions is meaningful. Recent reports
tant platform chemical and can be used as food additive, drug, have reported a series of acid catalysts that may be eective in
solvent and the precursor of methyl-tetrahydrofuran, which cellulose hydrolysis, such as metal salts, sulfonated carbon,
is used as a very popular fuel additive currently being (layered transition) metal oxides, heteropoly acid, sulfonated
investigated.79 silica nano/meso composites, niobium phosphate, and so
on.2146 However, in most of these studies, sugars or water-
soluble oligosaccharides, but not cellulose, were used as sub-
strates for the production of LA.
Key Lab for Advanced Materials, Research Institute of Industrial Catalysis East China
University of Science and Technology, # 130, Meilong Road, Shanghai, 200237, Previous studies showed that the type of acid (Brnsted and
P. R. China. E-mail: gzhlu@ecust.edu.cn, wangyanqin@ecust.edu.cn Lewis acids, as well their ratio), acid amount and acid strength

3846 | Green Chem., 2014, 16, 38463853 This journal is The Royal Society of Chemistry 2014
View Article Online

Green Chemistry Paper

concentration, similar to the results obtained using glucose


and HMF as the substrates.15,16 V. Ordomsky et al.44 compared
the performances of a series of solid acid catalysts and found
that higher acid amount leads to higher LA selectivity when
using solid acids as catalysts.
Subsequently, the hydrogenation of LA to GVL is convenient
and 100% yield of GVL, even on the industrial scale, can be
easily achieved in the presence of the Ru/C catalyst. S. G.
Published on 19 June 2014. Downloaded by Federal University of Minas Gerais on 24/04/2015 15:03:26.

Wettstein et al.9 have reported that a high GVL yield of 75%


was achieved from cellulose in a waterGVL biphasic system
using HCl and RuSn/C as hydrolysis and hydrogenation cata-
lysts, respectively. This approach for cellulose conversion elimi-
nates the need to separate the final product from the solvent,
because the GVL product is the solvent.
In our previous studies, mesoporous niobium phosphate
exhibited excellent performances in the conversion of fructose,
Scheme 1 The reaction pathway of converting cellulose to levulinic xylose, and even cellulose in pure water and a biphasic system
acid and GVL in water.
under mild conditions because of its suitable Lewis
and Brnsted acidity, as well as its hot-water-tolerant prop-
erty.33,38,47,48 Herein, we continue our study to investigate the
in solid catalysts had an influence on the selectivity to LA from catalytic performance of mesoporous niobium phosphate in
cellulose or glucose conversion. F. Chambon et al.25 investi- the direct conversion of cellulose to LA and the influence of
gated the performance of a series of heterogeneous solid acid modification of the acidic properties by doping with alumi-
catalysts in the hydrolysis of cellulose in hot water and found num on the selectivity of LA. Besides, the direct upgrading of
that the catalyst containing much more Lewis acid, such as the generated LA in the aqueous solution into GVL is investi-
ZrW and AlW, had higher cellulose hydrolysis level, but the gated as well.
main product was lactic acid. The highest LA yield was
achieved over sulfated zirconia (ZrS), which has both Brnsted
and Lewis acid sites. These results indicate that large amounts Experimental section
of solid Lewis acids can promote cellulose depolymerisation,
Chemicals
and Brnsted acid would enhance LA production.
V. Choudhary et al.39 also found that Lewis acid, which is All chemicals were purchased from Sinopharm Chemical
CrCl3 in their research, would speed up the isomerisation of Reagent Co. Ltd, except cellulose and the Nb precursor,
glucose, but its selectivity to HMF and LA was low when used niobium tartrate; the former was bought from Fluka Analytical
by itself. Thus, the interplay of Lewis and Brnsted acids is Co. Ltd and the latter was prepared in our laboratory according
important for the conversion of cellulose to LA, and excellent to the literature.32,33 All purchased chemicals were of analytical
yield may be obtained by using a solid acid with suitable Lewis grade and used without further purification.
and Brnsted acidity.
R. Weingarten et al.42 deeply investigated the selectivity of Catalysts preparation
HMF and LA with solid metal(IV) phosphate catalysts and Mesoporous niobium phosphate was synthesized according to
liquid acids by controlling the acid concentration and glucose our previous report.33 Typically, 0.01 mol of diammonium
conversion in the same level, and adjusting the molar ratio of hydrogen phosphate was dissolved in 20 mL water and
B/L acid from 0 to 1. It was also found that Lewis acid would adjusted to pH 2 using phosphoric acid; then 20 mL of 0.5 M
speed up the conversion of glucose, while the selectivity to niobium tartrate ( pH = 2) was added to the above solution
HMF and LA is low. The LA selectivity had an obvious with stirring. The mixed solution was slowly poured into
enhancement when the molar ratio of B/L acid was adjusted the previously prepared aqueous solution of cetyltrimethyl
from 1 : 3 to 1 : 1 using homogeneous or solid acid catalysts. ammonium bromide (CTAB), which contains 1.0 g of CTAB in
These results are similar to those of F. Chambon et al.,25 13 mL water, and the final pH was around 2. A large amount
giving us the enlightenment of designing the solid acid cata- of niobium-phosphorus precipitation appeared immediately
lyst with strong acid sites and a suitable B/L acid molar ratio. from the solution, and the mixture was stirred for an
In another method, the kinetic model of cellulose hydro- additional 60 min at 35 C, and then transferred into a Teflon-
lysis to LA was studied using hydrochloric acid by J. C. Shen lined autoclave and aged at 160 C for 24 h. The solid was fil-
et al.14 and using sulfuric acid by B. Girisuta et al.13 Both of tered after cooling down, washed with distilled water and
the research studies made the conclusion that the reaction of dried at 100 C overnight. Finally, the catalyst was obtained by
cellulose hydrolysis needs severe conditions such as higher calcination at 500 C for 5 h in air to remove organics. The syn-
temperature, higher acid concentration or lower substrate thesis of aluminium-modified porous niobium phosphate was

This journal is The Royal Society of Chemistry 2014 Green Chem., 2014, 16, 38463853 | 3847
View Article Online

Paper Green Chemistry

similar to the above procedure except for the addition of a cal- stirring. N2 was used as a protective gas and kept at a pressure
culated amount of aluminium precursor, which was prepared of about 0.8 MPa. The zero time was taken when the Teflon-
by dissolving aluminium hydroxide in 10 mL of 1.5 M oxalic lined stainless steel autoclave was placed in the heating
acid solution. mantle under magnetic stirring, which had already been
heated to the set temperature. After reaction, the catalyst was
Characterizations collected by centrifugation and washed with water and ethanol
several times, and then dried at 100 C for 24 hours for the
Nitrogen adsorptiondesorption isotherms of the catalysts
next run. The supernatant liquid in the centrifugation tube
Published on 19 June 2014. Downloaded by Federal University of Minas Gerais on 24/04/2015 15:03:26.

were measured at 96 C on a NOVA 4200e analyzer (Quanta-


was analyzed by an Agilent 1200 Series HPLC based on the
chrome Co. Ltd). Before the measurements, all samples were
external standard method. All analytical data were the averages
outgassed at 180 C for 12 h under vacuum to remove moisture
from three repeated reactions, and the error was less than 1%
and volatile impurities. The BET method was used to calculate
compared with the average data. The soluble polymer in the
the specific surface area, the pore size distribution curve was
liquid was separated by evaporation of water at 100 C, and the
calculated by the BarrettJoynerHalenda (BJH) method
insoluble humins were obtained by filtering. Together with the
through the desorption branch of the isotherm and the total
unused and used catalysts, the soluble polymer and insoluble
pore volume (Vt) was estimated at a relative pressure of 0.95.
humins were weighed and analyzed by EDS to determine the
The pyridine adsorption infrared (IR) spectra were recorded
element content.
on a Nicolet NEXUS 670 FT-IR spectrometer, with 32 scans at
In the conversion of cellulose to GVL, the mixture of 0.4 g
an eective resolution of 4 cm1. 30 mg of the catalyst was
solid acid catalyst, 0.04 g Ru/C (5%), 0.5 g cellulose and 10 mL
pressed into a self-supporting disk and placed in an IR cell
water was loaded into the Teflon-lined stainless steel autoclave
attached to a closed glass-circulation system. The catalyst disk
and reacted for 24 h as above; then it was quenched in cool
was dehydrated by heating at 400 C under vacuum in order to
water and N2 gas in the autoclave was replaced with H2 gas
remove physisorbed moisture. The IR spectrum background
and the reaction was restarted. The analysis of the product
was recorded at room temperature when the cell cooled down.
from the hydrogenation reaction was carried out with the
Pyridine vapor was then introduced into the cell at room temp-
same HPLC.
erature until equilibrium was reached. Subsequent evacuations
The analysis of the reaction products was carried out in a
were performed at 100, 200, 300 and 400 C for 60 min fol-
HPLC apparatus (Agilent 1200 Series) equipped with an ion-
lowed by spectral acquisitions at room temperature using the
exclusion column Aminex HPX-87H (Bio-Rad), eluting with an
background recorded before. The quantification of acid sites
aqueous solution of sulphuric acid as the mobile phase. The
was performed using the same expressions as those in the
LA and other products were analyzed using a refractive index
research article of Angela S. Rocha:49
 2 detector (Agilent G1362A) in the HPLC.
r For the calculation of the carbon balance, humins and
CL K L A1450 A1450
IMECL w other solid carbon-based compounds were separated by cen-
trifugation (together with solid catalyst) and then analyzed by
and
TG and EDX; for the part of soluble polymers, they were
 2
r obtained by evaporating water, formic acid and LA and then
CB K B A1540 A1540
IMECB w analyzed for carbon by EDX. The analysis of humins and
soluble polymers may not be accurate, but still gives infor-
mation for carbon distribution.
CL and CB are the concentrations of Lewis acid sites and
Brnsted acid sites in mol g1; A1450 and A1540 are the inte-
grated areas of bands at 1450 and 1540 cm1 in the original
data of FTIR spectra, as shown in the Results and discussion Results and discussion
section. KL and KB are molar extinction constants for Lewis
In our previous work, mesoporous niobium phosphate was
and Brnsted acid sites; IMECL and IMECB are integration
synthesized with high surface area and narrow pore size distri-
molar extinction coecients, 2.22 and 1.67 cm mol1 for
bution, and exhibited excellent activities in the conversion of
Lewis and Brnsted acids, respectively; r is the wafer radius in
fructose and xylose in pure water under mild conditions
cm and w is the wafer weight in g of the self-supporting cata-
because of the higher acid amount, co-existence of Lewis and
lyst disk.
Brnsted acidity, as well as its hot-water-tolerant pro-
perties.33,38,47 The total amount of acid was around 1000 mol
Catalytic reactions g1 and the molar ratio of B/L acid was ca. 1.8 at 373 K from
In the conversion of cellulose to LA, a batch reactor was used. Py-FTIR, the latter of which was much higher than those of AlW,
Typically, a mixture of 0.4 g catalyst, 0.5 g cellulose and 10 mL ZrW, ZrP, and SnP (B/L ratio normally lower than 1.0).25,42
water was loaded into a Teflon-lined stainless steel autoclave Meanwhile, LA was always co-produced in these reactions
and placed into a temperature-controlled electric heating although with low selectivity. Generally, high acid concen-
mantle with a thermocouple probed detector and magnetic trations and reaction temperatures can accelerate the reaction

3848 | Green Chem., 2014, 16, 38463853 This journal is The Royal Society of Chemistry 2014
View Article Online

Green Chemistry Paper

rate and enhance LA yield.13,14 Hence, in this work, the reac- Table 1 Chemical composition and BET properties of porous niobium
tion temperature was increased to 180 C, and ca. 40% yield of catalyst synthesized with dierent amounts of aluminium
LA was obtained after 24 h from 5 wt% cellulose solution cata-
SBET/ Pore size/
lyzed by mesoporous niobium phosphate. This yield of LA Catalyst m2 g1 nm Nb/at% Al/at% P/at%
was comparable with that of 38.6% using 0.9 mol L1 HCl as a
catalyst under 160 C from ca. 0.8 wt% cellulose solution14 and H-ZSM-5 318 0.55
NbOPO4 (A) 233 3.8 20.94 20.95
higher than those reported by R. Weingarten et al.,42 in which Al-NbOPO4 (B) 145 7.6 21.01 1.12 21.39
zirconium (ZrP) and tin (SnP) phosphates, or other solid acid Al-NbOPO4 (C) 102 7.3 18.98 2.12 20.82
Published on 19 June 2014. Downloaded by Federal University of Minas Gerais on 24/04/2015 15:03:26.

catalysts such as metal oxides and heteropoly acid containing Al-NbOPO4 (D) 50 9.6 17.45 2.49 18.61
Al, W, Zr, Cs and Sn were used as catalysts for glucose conver-
sion to LA under 160 C.25 Such an excellent performance may
be attributed to the unique Brnsted and Lewis acid property
of niobium phosphate.
Considering pure niobium phosphate with a high B/L
molar ratio (ca. 1.8), we would like doping of aluminium in
niobium phosphate to enhance Lewis acid because aluminium is
always considered as a strong Lewis acid site, and samples with
dierent amounts of aluminium were prepared using the same
method. The samples doped with 5 and 10% aluminium were
referred to as samples B and C; as a comparison, pure
niobium phosphate was named A. Another catalyst D was also Fig. 2 Py-FTIR spectra of catalyst A (a) and catalyst D (b).
prepared by lowering the amount of CTAB (ca. a quarter of the
initial on catalyst D) during the hydrothermal synthesis
process. ICP analysis and N2 sorption were conducted to inves-
tigate the change of the composition and structural para- Table 2 The weak/medium/strong acid sites amounts of catalysts A, B,
meters after aluminium introduction and the results are C and D calculated from pyridine adsorption infrared (Py-FTIR) spectra
presented in Fig. 1 and Table 1, respectively. It is found that
Niobium phosphate A B C D
the molar ratios of Nb/P in dierent samples were nearly the
same, close to 1, while the surface area decreased gradually Brnsted acid/ Weak 373473 K 113.0 127.9 109.3 122.4
with the doping of aluminium. The acidic properties (type, mol g1 Medium 473673 K 397.6 268.0 253.4 136.1
Strong >673 K 130.9 215.2 323.7 234.1
strength and amount) were also studied through pyridine Total 641.5 611.1 686.4 492.6
adsorbed infrared spectra (Py-FTIR) and the data are collected Lewis acid/ Weak 373473 K 68.2 57.9 51.7 31.2
in Fig. 2 and Table 2. It is found that the strong Brnsted and mol g1 Medium 473673 K 231.3 201.2 75.2 119.4
Strong >673 K 45.6 112.7 296.9 318.7
Lewis acid sites increased with increasing the amount of Al, Total 345.1 371.8 423.8 469.3
especially the strong Lewis acid sites, which increased from
45.6 to 296.9 mol g1, indicating that aluminium is indeed a
good strong Lewis acid centre, while the medium acid sites fer from a medium acid site to a strong acid site, which is
decreased with increasing the Al doping, maybe due to a trans- unclear till now.
The catalytic performances of various catalysts including
H-ZSM-5, a strong solid acid with a molar ratio of Si/Al = 25, in
the conversion of cellulose to LA are presented in Fig. 3. It can
be seen that the LA yield increased with the increase of alu-
minium amount in the catalyst, but the conversions of cellu-
lose are similar, all above 90%. This kind of tendency may
imply that the increase of strong Brnsted and Lewis acids by
aluminium doping will promote the generation of LA from
cellulose and the highest LA yield of 52.9% can be obtained by
using catalyst D, which contains 2.49% aluminium. This value
is 12% higher than that from pure mesoporous niobium phos-
phate as a catalyst, and 24% higher than that from H-ZSM-5.
Even though such a value of LA yield is still lower than that of
6070% from liquid acids (e.g., HCl) in the biphasic phase,9 it
is already much higher than those obtained from any other
kind of solid acid catalyst investigated so far. The analytical
data were the average from three repeated reactions, and the
Fig. 1 N2 adsorption/desorption isotherms of catalysts A, B, C and D. error was less than 1% compared with the average data.

This journal is The Royal Society of Chemistry 2014 Green Chem., 2014, 16, 38463853 | 3849
View Article Online

Paper Green Chemistry

Table 3 Chemical composition and BET properties of used catalyst D


and humins

SBET/ Pore
m2 g1 size/nm Nb/at% Al/at% P/at%

Recycled catalyst D 39 8.4 18.98 1.29 21.54

C/at% H/at% O/at%


Published on 19 June 2014. Downloaded by Federal University of Minas Gerais on 24/04/2015 15:03:26.

Humins 59.18 8.26 32.10


Soluble polymers 43.76 16.42 38.94

carbon in humins and soluble polymers are ca. 60% and 44%,
respectively, the former of which was separated and analyzed
by EDS and are shown in Table 3, and the latter was calculated
from the structure of humins41 and monomers of oligo-
saccharides. In addition, no lactic acid, HMF, glucose or other
small organic molecules were detected in the reaction solu-
Fig. 3 Cellulose conversion and LA yield from catalysts A, B, C, D, and
H-ZSM-5 (Al/Si = 1/25). Reaction conditions: 0.5 g of cellulose, 0.4 g of tion, and the main by-products are humins and soluble poly-
catalyst, 10 g of H2O, 180 C, 24 h. Calculated from the conversion of saccharides. According to previous studies,25,39 it is believed
anhydroglucose unit in cellulose. that small molecules, such as lactic acid would appear only
when there is not enough Brnsted acid in the reaction
system. Once again, the importance of the bonding between
Brnsted and Lewis acids is emphasized. It is also notable that
in the products distribution catalyzed by these four niobium-
based catalysts, the percentage of soluble oligomers and poly-
mers remains stable, and the increase of FA and LA would
result from the decrease of humins.
To investigate the influence of strong acidity, especially
strong Lewis acids on the selectivity/yield of LA, the LA selecti-
vity/yield from glucose and HMF catalyzed by these four cata-
lysts were investigated and compared with cellulose conversion
and presented in Fig. 5. In all cases, the conversion of all sub-
strates is above 95%. It is found that the LA yield from HMF
had the same tendency as that from cellulose, indicating that
the solid acid catalyst with more aluminium content, that is,
stronger Lewis acid sites could increase the LA yield. The
strong Lewis acid would enhance the selectivity to LA in HMF
conversion was never mentioned before and will be further
confirmed by using a strong liquid Lewis acid later. The result
from glucose has the same tendency, too, but conflicted with
the previous reports,25,39,43 especially the studies reported by
V. V. Ordomsky et al.,43 where the high molar ratio of B/L acid
would promote the formation of LA from glucose. This surpris-
Fig. 4 The distributions of carbon elements in the products by using
ing result may be due to the high B/L molar ratio (ca. 1.21.8)
various acid catalysts under the same conditions as in Fig. 3. Calculated
from the total number of carbon atoms. in niobium phosphate, much higher than those other reported
catalysts (lower than 1.0), so the addition of aluminium
increased the strong Lewis acid and thus enhanced the selecti-
The distribution of carbon element in the products was vity to LA from HMF. It means that there is a balance of Lewis
analyzed by HPLC and thermal analysis, and the result is pre- and Brnsted acids for the conversion of HMF/glucose/cellu-
sented in Fig. 4. Cellulose conversion is above 95% when lose to LA, not the higher B/L ratio, a better selectivity to LA.
using a niobium-based catalyst and 93% for H-ZSM-5, indicat- On the other hand, the final LA yield from glucose is lower
ing that cellulose can be fully converted using these solid acid than that from cellulose with all four catalysts. This may be
catalysts. The ratio of carbon atoms in formic acid and levuli- due to the side reactions during the conversion of glucose to
nic acid analyzed by HPLC is always 1 : 5 in solution (of equal HMF, which happens much easier at a high concentration of
molecule amounts), which is consistent with the report that glucose. While in the conversion of cellulose, glucose just
LA and formic acid are equimolarly produced. The contents of existed as an intermediate with very low concentration, which

3850 | Green Chem., 2014, 16, 38463853 This journal is The Royal Society of Chemistry 2014
View Article Online

Green Chemistry Paper


Published on 19 June 2014. Downloaded by Federal University of Minas Gerais on 24/04/2015 15:03:26.

Fig. 6 LA generation from HMF catalyzed by dierent catalysts under


Fig. 5 Comparison of LA production from dierent substrates catalyzed
the conditions of 180 C, 24 h. The total acid amount was controlled at
by four catalysts, under the same reaction conditions as in Fig. 3.
0.0004 mol in 10 mL of water.

prevents the formation of humins or soluble polymers and


therefore leads to high LA yield.
It is also worth noting that the pH value in the system
before the reaction started was 6.1 (measured after 2 h of stir-
ring) while the final LA solution was around 1.8, which means
that the continuing increase of the Brnsted acid ratio in the
system was caused by the dissociation of LA (COOH) in water,
and the acid catalyst was stable in water.
The enhancement of a strong Lewis acid to the selectivity of
LA was further confirmed by the addition of a liquid Lewis
acid, lanthanum trifluoroacetate [(TfO)3La] in the catalytic
system. The pH value of (TfO)3La in a 0.01 mol L1 aqueous
solution was 5.42, close to that of distilled water, meaning that
it remained as a solid Lewis acid in water and almost no H+
ions generated through the hydrolysis of this salt. Three cata-
lytic systems were compared in HMF conversion, with and Fig. 7 Products distribution in various reaction temperatures. Con-
without (TfO)3La, including homogeneous acids, HCl, trifluoro- ditions: cellulose 0.5 g, catalyst D 0.4 g, H2O 10 g, reaction time 24 h.
acetic acid [(TfO)3H] and a solid catalyst with a high B/L ratio,
mesoporous niobium phosphate (catalyst A), the results are
presented in Fig. 6. All the three catalysts showed an enhance- yield was obtained, up to 52.9%. Further increasing tempera-
ment in LA yield when (TfO)3La was added to the aqueous ture to 200 C, the LA yield decreased, which may be attributed
solution, indicating an addition of Lewis acid can enhance the to the speeding up of the side reactions and lead to the
selectivity of LA from HMF although the mechanism is formation of more solid humins.
unclear. All these results explained the enhancement of LA The reusability of catalyst D is also investigated and pre-
yield from cellulose with aluminium-doped niobium phos- sented in Fig. 8. The conversion of cellulose decreased gradu-
phate as a catalyst. ally, while the selectivity to LA remained stable except for the
The influence of reaction temperature on products distri- second time, which may be due to the rearrangement of the
bution was also investigated to optimize the reaction condition acid sites on the solid acid catalysts in aqueous solution,
and is presented in Fig. 7. Clearly, a temperature higher than which is the common characteristic for most of the solid acid
140 C is needed for the production of LA, because at this catalysts.33 The gradual decrease of the catalytic activity may
temperature, a high amount of unreacted cellulose (ca. 40%) be caused by the deposition of humins on the surface of the
and unhydrolyzed water-soluble compounds were present as catalyst or the leaching of Al, Nb or P elements. To confirm the
shown in the products distribution. Other intermediates, influence of humins deposition, the used catalyst was treated
including glucose, HMF and oligosaccharide also existed. At with hot methanol and found that a certain degree of catalytic
160 C, more LA was produced with less cellulose left. When activity was really recovered, indicating the influence of
the temperature was further increased to 180 C, higher LA humins deposition. Furthermore, the used catalyst was

This journal is The Royal Society of Chemistry 2014 Green Chem., 2014, 16, 38463853 | 3851
View Article Online

Paper Green Chemistry

pressure higher than 3 MPa to make the reaction going fast


enough. It is also interesting to find that the yield of GVL is
higher than that of LA (56.9% vs. 52.9%), which may be due to
the strong adsorption of LA on the surface of the catalyst,
which reduced concentration of LA in solution; while during
hydrogenation, the adsorbed LA turned into GVL, and then
was released back into water. To verify this assumption, 0.4 g
catalyst D, 10 g water and 0.215 g LA (0.185 mol L1, as high as
Published on 19 June 2014. Downloaded by Federal University of Minas Gerais on 24/04/2015 15:03:26.

60% LA yield from 0.5 g cellulose) were loaded in a reactor and


kept at 180 C for 24 h, then cooled down and analyzed by
HPLC and GC-MS. It is found that the amount of LA dropped
to 0.174 mol L1 in solution, while no by-products were
detected, indicating the adsorption of LA. When replacing LA
with the same amount of GVL, the yield of GVL dropped from
0.185 mol L1 to 0.183 mol L1, much lower than that of LA.
These results confirmed the assumption of the adsorption of
Fig. 8 Catalyst reusability. Reaction conditions: 0.5 g of cellulose, 0.4 g LA on the catalyst, and once converted to GVL by hydrogen-
of catalyst D, 10 g of H2O, 180 C, 24 h. ation, most of them would be released back into the solution.
Thus, the yield of GVL looks higher than that of LA.

collected and analyzed by thermal gravimetric analysis,


Py-FTIR, EDS and nitrogen sorption. The BET surface area of
the recycled and re-calcined catalyst D was around 39 m2 g1, Conclusions
smaller than the fresh one (50 m2 g1), while the pore size did
A high yield of 52.9% of LA was achieved over Al-doped meso-
not change obviously, compared with the fresh catalyst D. EDS
porous niobium phosphate in aqueous solution, which is
analysis showed that the contents of Nb and P were almost the
better than any other solid acid catalysts reported so far. The
same as the fresh one, but Al showed an obvious loss. Py-FTIR
high selectivity and yield of LA were attributed to the strong
of used catalyst D also showed the decrease of the total
acid strength and a suitable B/L acid molar ratio (1.2 : 1) of
acid amount, especially the decrease of strong acids. Thus, the
this catalyst. The doping of aluminium in niobium phosphate
leaching of Al could be another reason for the decrease of
enhanced the strong Lewis and Brnsted acids, especially the
cellulose conversion. In the future, we would enhance the cata-
strong Lewis acid and led to the increase in LA yield from
lyst stability by using the new synthesis method or incorporate
cellulose as well as from glucose and HMF. The addition of
other Lewis acid sites into NbOPO4.
(TfO)3La into the catalytic system containing another catalyst
The production of GVL comes from the hydrogenation of
(HCl, (TfO)3H, niobium phosphate) further confirmed a suit-
LA (Fig. 9) after cellulose hydrolysis for 24 h without separ-
able ratio of Brnsted/Lewis acid is important for higher
ation. It can be found that hydrogenation of LA needs a H2
selectivity of LA from HMF, as well as from cellulose. After pro-
duction of LA in aqueous solution of cellulose, it can be com-
pletely converted to -valerolactone through hydrogenation
over the Ru/C catalyst without further separation.

Acknowledgements
This project was supported financially by the NSFC of China
(no. 21101063 and 21273071), the Science and Technology
Commission of Shanghai Municipality (13520711400,
13JC1401902, 10dz2220500) and the Fundamental Research
Funds for the Central Universities, China.

Notes and references


1 A. Corma, S. Iborra and A. Velty, Chem. Rev., 2007, 107,
Fig. 9 Hydrogenation of LA generated through cellulose conversion,
2411.
catalysed by 0.04 g of Ru/C (5 wt% Ru) under dierent H2 pressures, 2 G. W. Huber, S. Iborra and A. Corma, Chem. Rev., 2006,
180 C, 12 h. 106, 4044.

3852 | Green Chem., 2014, 16, 38463853 This journal is The Royal Society of Chemistry 2014
View Article Online

Green Chemistry Paper

3 S. N. Naik, V. V. Goud, P. K. Rout and A. K. Dalai, Renewable 27 J. Potvin, E. Sorlien, J. Hegner, B. DeBoef and B. L. Lucht,
Sustainable Energy Rev., 2010, 14, 578. Tetrahedron Lett., 2011, 52, 5891.
4 T. Werpy, G. Petersen, A. Aden, J. Bozell, J. Holladay, 28 P. Lanzafame, D. M. Temi, S. Perathoner, A. N. Spadaro
A. Manheim, D. Eliot, L. Lasure and S. Jones, Top Value and G. Centi, Catal. Today, 2013, 179, 178.
Added Chemicals from Biomass, U.S. Department of Energy, 29 R. Weingarten, W. C. Conner Jr. and G. W. Huber, Energy
Oak Ridge, TN, 2004, vol. 1. Environ. Sci., 2012, 5, 7559.
5 F. Rataboul and N. Essayem, Ind. Eng. Chem. Res., 2011, 50, 30 P. A. Son, S. Nishimura and K. Ebitani, React. Kinet. Mech.
799. Cat., 2013, 106, 185.
Published on 19 June 2014. Downloaded by Federal University of Minas Gerais on 24/04/2015 15:03:26.

6 E. I. Grbz, D. M. Alonso, J. Q. Bond and J. A. Dumesic, 31 A. Tiziana, B. Guido, C. Carlo, G. Mario, R. Galletti,
ChemSusChem, 2011, 4, 357. A. Maria and S. Glauco, J. Mol. Catal. A: Chem., 2000, 151,
7 H. Heeres, R. Handana, D. Chunai, C. B. Rasrendra, 233.
B. Girisuta and H. J. Heeres, Green Chem., 2009, 11, 1247. 32 A. Sarkar and P. Pramanik, Microporous Mesoporous Mater.,
8 I. T. Horvath, H. Mehdi, V. Fabos, L. Boda and L. T. Mika, 2009, 117, 580.
Green Chem., 2008, 10, 238. 33 Y. Zhang, J. J. Wang, J. W. Ren, X. H. Liu, X. C. Li, Y. J. Xia,
9 S. G. Wettstein, D. M. Alonso, Y. Chong and J. A. Dumesic, G. Z. Lu and Y. Q. Wang, Catal. Sci. Technol., 2012, 2,
Energy Environ. Sci., 2012, 5, 8199. 2485.
10 The B F Goodrich Company, Process for the Manufacture of 34 Z. Sun, M. X. Cheng, H. C. Li, T. Shi, M. J. Yuan,
Levulinic Acid and Ethers [P]. US 4236021, 1980. X. H. Wang and Z. J. Jiang, RSC Adv., 2012, 2, 9058.
11 J. Zhang, S. B. Wu, B. Li and H. D. Zhang, ChemCatChem, 35 J. J. Wang, J. W. Ren, X. H. Liu, J. X. Xi, Q. N. Xia, Y. H. Zu,
2012, 4, 1230. G. Z. Lu and Y. Q. Wang, Green Chem., 2012, 14, 2506.
12 S. V. de Vyver, J. Geboers, P. A. Jacobs and B. F. Sels, Chem- 36 W. J. Xu, H. F. Wang, X. H. Liu, J. W. Ren, Y. Q. Wang and
CatChem, 2011, 3, 82. G. Z. Lu, Chem. Commun., 2011, 7, 3924.
13 B. Girisuta, L. P. B. M. Janssen and H. J. Heeres, Ind. Eng. 37 J. J. Wang, W. J. Xu, J. W. Ren, X. H. Liu, G. Z. Lu and
Chem. Res., 2007, 46, 1696. Y. Q. Wang, Green Chem., 2011, 13, 2678.
14 J. C. Shen and C. E. Wyman, AIChE J., 2012, 58, 236. 38 J. X. Xi, Y. Zhang, Q. N. Xia, X. H. Liu, J. W. Ren, G. Z. Lu
15 B. Girisuta, L. P. B. M. Jansen and H. J. Heeres, Chem. Eng. and Y. Q. Wang, Appl. Catal., A, 2013, 459, 52.
Res. Des., 2006, 84(A5), 339. 39 V. Choudhary, S. H. Mushrif, C. Ho, A. Anderko,
16 B. Girisuta, L. P. B. M. Janssen and H. J. Heeres, Green V. Nikolakis, N. S. Marinkovic, A. I. Frenkel, St. I. Sandler
Chem., 2006, 8, 701. and D. G. Vlachos, J. Am. Chem. Soc., 2013, 135, 3997.
17 J. Y. Cha and M. A. Hanna, Ind. Crop. Prod., 2002, 16, 109. 40 F. F. Wang, C. L. Liu and W. S. Dong, Green Chem., 2013,
18 R. P. Swatloski, S. K. Spear, J. D. Holbrey and R. D. Rogers, 15, 2091.
J. Am. Chem. Soc., 2002, 124, 4974. 41 S. K. R. Patil and C. R. F. Lund, Energy Fuels, 2011, 25,
19 J. B. Binder and R. T. Raines, J. Am. Chem. Soc., 2009, 131, 4745.
1979. 42 R. Weingarten, Y. T. Kim, G. A. Tompsett, A. Fernndez,
20 S. Hu, Z. Zhang, Y. Zhou, B. Han, H. Fan, W. Li, J. Song K. S. Han, E. W. Hagaman, W. C. Conner, J. A. Dumesic
and Y. Xie, Green Chem., 2008, 10, 1280. and G. W. Huber, J. Catal., 2013, 304, 123.
21 R. Rinaldi, R. Palkovits and F. Schth, Angew. Chem., Int. 43 V. V. Ordomsky, V. L. Sushkevich, J. C. Schouten, J. van der
Ed., 2008, 47, 8047. Schaaf and T. A. Nijhuis, J. Catal., 2013, 300, 37.
22 S. Suganuma, K. Nakajima, M. Kitano, D. Yamaguchi, 44 V. V. Ordomsky, J. V. der Schaaf, J. C. Schouten and
H. Kato, S. Hayashi and M. Hara, J. Am. Chem. Soc., 2008, T. A. Nijhuis, ChemSusChem, 2012, 5, 1812.
130, 12787. 45 W. Zeng, D. G. Cheng, H. G. Zhang, F. Q. Chen and
23 A. Takagaki, C. Tagusagawa and K. Domen, Chem. X. L. Zhan, React. Kinet. Mech. Cat., 2011, 100, 377.
Commun., 2008, 5363. 46 H. F. Ren, Y. G. Zhou and L. Liu, Bioresour. Technol., 2013,
24 S. V. de Vyver, L. Peng, J. Geboers, H. Schepers, F. de 129, 616619.
Clippel, C. J. Gommes, B. Goderis, P. A. Jacobs and 47 X. C. Li, Y. Zhang, Y. J. Xia, B. C. Hu, L. Zhong, Y. Q. Wang
B. F. Sels, Green Chem., 2011, 12, 1560. and G. Z. Lu, Acta Phys.Chim. Sin., 2012, 28, 2349.
25 F. Chambon, F. Rataboul, C. Pinel, A. Cabiac, E. Guillon 48 I. Nowak and M. Ziolek, Chem. Rev., 1999, 99, 3603.
and N. Essayem, Appl. Catal., B, 2011, 105, 171. 49 A. S. Rocha, A. M. S. Forrester, M. H. C. de la Cruz, C. T. da
26 J. Hegner, K. C. Pereira, B. DeBoef and B. L. Lucht, Tetra- Silva and E. R. Lachter, Catal. Commun., 2008, 9,
hedron Lett., 2010, 51, 2356. 1959.

This journal is The Royal Society of Chemistry 2014 Green Chem., 2014, 16, 38463853 | 3853

You might also like