Sie sind auf Seite 1von 10

Catalysis

Science &
Technology
View Article Online
PAPER View Journal

High yield conversion of cellulosic biomass into


5-hydroxymethylfurfural and a study of the
Published on 24 May 2016. Downloaded by University College London on 27/05/2016 10:49:57.

Cite this: DOI: 10.1039/c6cy00820h


reaction kinetics of cellulose to HMF conversion in
a biphasic system
Luqman Atanda,a Muxina Konarova,a Qing Ma,a Swathi Mukundan,a
Abhijit Shrotrib and Jorge Beltramini*a

Catalytic technology for cellulosic biomass conversion has been proven to be a promising approach for
valuable chemical feedstock production. However, its recalcitrant nature is a major limitation to unlocking
the carbohydrate biopolymer content and their subsequent conversion into 5-hydroxymethylfural (HMF).
This paper investigates the production of HMF using glucose, cellulose, sugarcane bagasse and rice husk
as the feedstocks. Acid dehydration of the carbohydrate sources was conducted in a biphasic system of
waterMeTHF modified with N-methyl-2-pyrrolidone (NMP) over a phosphated TiO2 catalyst. The catalyst
displayed a very good catalytic performance for the conversion of glucose into HMF (91% yield). More so, it
is suitable for the selective conversion of mechanocatalytic depolymerized cellulose to 74.7% yield of HMF.
Cellulosic biomass could also be directly converted into HMF and furfural in reasonable yields. The effi-
Received 14th April 2016, ciency of biomass-to-HMF production was further advanced after biomass fractionation treatment. Re-
Accepted 17th May 2016
markable yields of 72% and 65% HMF were produced from sugarcane bagasse and rice husk, respectively.
DOI: 10.1039/c6cy00820h
Finally, the reaction kinetics of solubilized cellulose to HMF conversion was investigated and a simplified ki-
netic model comprising two reaction steps was developed: (1) hydrolysis of cello-oligomers to glucose and
www.rsc.org/catalysis (2) glucose dehydration to HMF.

1. Introduction HMF is a vital building block compound whose derivatives


can be applied in the production of solvents, polymers and
Alternatives to non-renewable sources of chemical and energy fuels.57 The main route extensively reported for obtaining
production have gained a lot of interest in recent years to HMF is by acid-catalyzed dehydration of monosaccharides,
meet the increasing global demand for these commodities. such as glucose and fructose.5,7,8 Currently, efficient produc-
The transition from a fossil-fuel-driven to a greener economy tion of HMF from cellulose or directly from lignocellulosic
may be achieved through the utilization of lignocellulosic bio- biomass using ionic liquid has recorded significant
mass, which is acknowledged as the world's most abundant progress.920 However, the commercial viability of this ap-
renewable carbon resource.1 As a result, considerable research proach is faced with challenges owing mainly to the high cost
is ongoing to advance catalytic technologies through novel of ILs and the prohibitive cost of separation.21 Adopting a bi-
catalyst design and innovative reaction engineering to realize phasic solution of waterorganic system is favoured as a
sustainable biorefinery processes for the production of highly effective and economically acceptable route of selec-
chemicals and fuels.24 A key step in this direction is the tively producing HMF on an industrial scale.2125 The contin-
transformation of the biopolymer carbohydrate components, uous partitioning of HMF into the organic phase prevents its
especially cellulose, which constitutes a major percentage (up re-polymerization or decomposition and thus improves the
to 50%) of biomass composition, to produce valuable plat- recovery efficiency of high-purity product. More importantly,
form chemicals like 5-hydroxymethylfurfural (HMF). realizing a low boiling point biomass-derived solvent with a
high partitioning coefficient will minimize the intensive re-
a
Nanomaterials Centre, Australian Institute for Bioengineering & Nanotechnology quirements for product recovery, thereby facilitating the over-
and School of Chemical Engineering, The University of Queensland, St. Lucia,
all industrial viability of carbohydrate dehydration in bi-
Brisbane, Queensland 4072, Australia. E-mail: j.beltramini@uq.edu.au
b
Institute for Catalysis, Hokkaido University, Kita 21 Nishi 10, Kita-Ku, Sapporo
phasic systems.
001-0021, Japan The design of an effective, easily separable and reusable
Electronic supplementary information (ESI) available. See DOI: 10.1039/c6cy00820h catalyst is also crucial for HMF synthesis. Metal chlorides in

This journal is The Royal Society of Chemistry 2016 Catal. Sci. Technol.
View Article Online

Paper Catalysis Science & Technology

tandem with mineral acids constitute the dominant catalyst furfural, titaniumIV) butoxide, n-butanol, ammonium hydroxide
system (homogeneous catalysis) used for acid dehydration of solution (28 wt%), ammonium phosphate monobasic,
carbohydrates in biphasic media.23,2629 Despite their effec- n-methyl-2-pyrrolidone, sodium hydroxide, barium hydroxide,
tiveness to produce high yields of HMF, they are still faced zinc chloride and aluminium chloride hexahydrate were
with the challenge of safe handling and corrosion hazards obtained from Sigma-Aldrich. 2-Methyltetrahydrofuran and
for large-scale processing. Meanwhile, solid acids may poten- hydrochloric acid (32 wt%) were purchased from Merck
tially overcome the above-mentioned limitations, though they Millipore. TinIV) chloride pentahydrate and copperII) chlo-
present additional challenges of low product yield compared ride dihydrate were both supplied by Ajax Finechem. De-
to homogeneous catalysts and sometimes require frequent re- pithed sugarcane bagasse was provided by Sugar Research
Published on 24 May 2016. Downloaded by University College London on 27/05/2016 10:49:57.

generation due to deactivation arising from partial/complete Institute, Queensland University of Technology, and rice husk
active site blockage by deposited humin compounds. Conse- was obtained from local suppliers. All the chemicals were
quently, the emphasis is to design an effective single catalyst used without further purification. Ultrapure water (18.2 M
system that is recalcitrant to humin formation. Zeolites,30,31 cm1) from the ELGA distillation system was used for all
heteropoly acids,3234 sulfonated polymers,35 phosphates,36,37 the experiments.
metal oxides38 and mesoporous metal phosphates39,40 are
some of the solid acids reported to have good catalytic perfor- 2.2. Catalyst preparation
mance for biphasic dehydration of carbohydrates. Recently, Neutral amine solgel technique was used to prepare
we reported the development of phosphated TiO2 by a two- phosphated TiO2 nanoparticles using titaniumIV) butoxide as
step synthesis method as a highly active and selective catalyst the TiO2 precursor. The sol was prepared by dropwise addi-
for the conversion of carbohydrates to HMF.41,42 The excel- tion of the alkoxide precursor into an aqueous solution of
lent catalytic behavior of phosphated TiO2 was rationalized in ammonium phosphate monobasic containing n-butanol un-
terms of its TiOP bonding, resulting in an acid bifunctional der stirring. The mole ratio of butanol/alkoxide/ammonium
nanoparticle with high surface acidity.41 The reaction me- phosphate/water in the initial mixture is 1/0.125/0.0124/2. Fi-
dium was found to play a complementary role in favouring nally, the pH of the solution was controlled and maintained
the formation of the desired product. By using a waterTHF at 7 by adding ammonium hydroxide solution and the mix-
biphasic system modified with NMP, the catalytic perfor- ture was maintained under reflux at 353 K for another 24 h.
mance of the catalyst was remarkably enhanced.42 After gelation, excess solvent was evaporated under reduced
Herein, we attempted a more facile catalyst preparation pressure and the sample was dried at 353 K overnight. Then,
technique by synthesizing phosphated TiO2 via a one-pot the resultant solid was calcined at 873 K for 4 h.
route. The choice of phosphated TiO2 catalyst in the present
study is based on the fact that it provided a balanced 2.3. Catalyst characterization
Brnsted/Lewis acidity required for the effective transforma-
tion of glucose-based carbohydrates.42 The structure and po- XRD patterns were obtained using a Rigaku Miniflex diffrac-
rosity characteristics of the as-synthesized material were veri- tometer equipped with monochromatic CoK radiation
fied by employing XRD and N2 porosimetry analyses. (30 kV, 15 mA). Analyses were performed at a scan speed of
Microscopic analysis including field emission scanning 5 min1 with a step of 0.02 in the 2 range of 1090. N2
electron microscopy (FE-SEM) and high-resolution transmis- adsorptiondesorption isotherms were obtained at 77 K on a
sion electron microscopy (HR-TEM) were applied to examine Micromeritics Tristar II 3020 analyzer surface area and
the morphology of the material. XPS, EDX and ICP analyses porosity analyzer. The sample was outgassed overnight at
were used to examine the chemical composition of the 473 K before measurement. SEM and TEM micrographs were
resulting material. The nanoparticle was tested for the cata- obtained with JEOL 7100 and JEOL 2100 electron micro-
lytic conversion of carbohydrates to HMF in a waterMeTHF scopes, respectively. Field emission scanning electron
biphasic medium. The reaction pathway towards HMF forma- microscopy (FE-SEM) and high resolution transmission
tion was examined, which was monitored through time analy- electron microscopy (HR-TEM) studies were conducted on a
sis of cellulose, cellobiose and glucose conversion reactions. JEOL JSM-7100 at 20 kV and a JEOL JEM 2010 microscope at
Based on the observed data, we developed a simplified reac- 200 kV, respectively. EDS was used as a chemical microanaly-
tion model and performed kinetic studies to obtain impor- sis technique to characterize the elemental composition of
tant reaction parameters. Direct utilization of raw biomass the nanomaterial, whereas the quantitative analysis of tita-
feedstocks such as sugarcane bagasse and rice husk was also nium and phosphorus contents was determined by induc-
examined for the production of HMF. tively coupled plasma (ICP) analysis carried out in a Mile-
stone Ethos-1 microwave digester using a HNO3 : HCl : HF
acid ratio of 1 : 3 : 1.
2. Experimental
2.1. Chemicals and materials 2.4. Two-stage biomass fractionation
Glucose, fructose, cellobiose, Sigmacell 20 m micro-crystalline Two-stage fractionation of rice husk and sugarcane bagasse
cellulose (MCC), oxalic acid, 5-hydroxymethylfurfural (HMF), was carried out by alkali pulping followed by the OrganoCat

Catal. Sci. Technol. This journal is The Royal Society of Chemistry 2016
View Article Online

Catalysis Science & Technology Paper

process. Alkali treatment of biomass was carried out with


2.5% NaOH at 353 K for 2 h in a solid/liquid ratio of 1 : 10 for
rice husk and 1 : 15 for sugarcane bagasse. Subsequently, the
alkali-treated materials were subjected to the OrganoCat pro- where nC6 and nCo6 denote the number of moles of C6 sugar
cess in a Parr reactor vessel following the procedure reported in the product and feed, respectively, and ni is the number of
by Grande et al.43 10 g of rice husk was suspended in a solu- moles of identified products (HMF, furfural etc.)
tion of 100 mL of water (150 mL of water is used in the case
of sugarcane bagasse) and oxalic acid (0.1 M), and 50 mL of 3. Results and discussion
MeTHF was added into the vessel. Thereafter, the tempera-
3.1. Catalyst characterization
Published on 24 May 2016. Downloaded by University College London on 27/05/2016 10:49:57.

ture was ramped to and held at 413 K for 3 h under 10 bar Ar


pressure with vigorous stirring. The XRD pattern of the material is shown in Fig. 1a. The
presence of the anatase phase of titania is confirmed by ob-
2.5. Mechanocatalytic depolymerization serving 2 diffraction peaks at 28, 45, 57 and 74, which corre-
Mechanocatalytic depolymerization of cellulose/biomass was spond to the (101), (004), (200) and (204) planes, respectively.
conducted by ball milling of acidulated substrate sample The position and intensity of these peaks are in agreement
using a procedure described elsewhere.44 In a typical method, with JCPDS File No. 21-1272. The broadening of the diffrac-
H2SO4 (2.5 mmol) was diluted to a volume of 40 mL. tion peaks indicates small-sized nanocrystals, and by using
Sigmacell microcrystalline cellulose/biomass (10 g) was then Scherrer's equation, the average crystallite size is calculated
added to this solution at room temperature, and the solution as 7.7 nm from the diffraction plane (101).
was stirred for a few minutes. The resulting slurry was dried The nitrogen adsorptiondesorption isotherm shown in
under reduced pressure, followed by overnight air drying at Fig. 1b is representative of a type IV isotherm with onset cap-
323 K. The acidulated cellulose/biomass thus obtained was illary condensation at P/P0 = 0.6, typical of a mesoporous par-
then milled in a planetary ball mill using 5 mm stainless ticle. The specific surface area of the particle is 125 m2 g1,
steel balls, with a substrate-to-ball weight ratio of 1 : 10. The while the pore diameter is estimated to be 6.04 nm.
mill was operated at 400 rpm, with a 15 min pause after every XPS analysis was carried out to analyse the chemical com-
30 min of continuous milling for a total milling time of 10 h. position and elucidate the chemical state of the nanomaterial
The milling time reported refers only to the active milling as shown in Fig. 2a.
time. The pause allowed dissipation of heat generated during The high resolution Ti 2p spectrum exhibits two peaks at
milling, which prevented overheating of reactants. 459 and 465, which are characteristic of Ti4+. P 2p spectra
could be observed at a binding energy of 134.1 eV, which in-
2.6. Catalytic test and product analysis dicates that phosphorus exits in the pentavalent oxidation
state (P5+). Deconvolution of the O 1s region by peak fitting
2.6.1 Catalytic test. Carbohydrate conversion reaction was revealed the presence of three species of oxygen at 530.2,
carried out in a stainless steel reaction vessel (Parr Instru- 531.1 and 533.4 eV corresponding to TiO, PO and OH
ment). The mixture of carbohydrate substrate, water, MeTHF, bonding, respectively.41,45
NMP and catalyst was charged into the reactor vessel and The SEM micrograph (Fig. 3a) shows that the sample
sealed. The system was purged with high purity (99.9%) Ar consisted of plate-like clusters of size 2040 m, and based
gas and then pressurized to 20 bar. The reactor temperature on HRTEM data (Fig. 3b), these particles consist of anatase-
was raised to the reaction temperature, kept under vigorous TiO2 nanocrystals having a mean size of 8.8 nm, which agrees
stirring and held at the set temperature for a predetermined well with the result obtained from XRD. In addition, the EDS
reaction time. Each experimental run was repeated three spectrum (Fig. S1) confirmed that the nanoparticle is
times with a typical error of 2%. elementally composed of Ti, O and P, with the C peak origi-
2.6.2 Product analysis. The products were analyzed using nating from carbon sputtering prior to SEM analysis, and the
a Shimadzu Prominence HPLC system equipped with an ana-
lytical column (Bio-Rad Aminex HPX-87H) and both RID-10
(refractive index) and SPD-M20 A (UV-Vis) detectors. HPLC
was operated under the following conditions: oven tempera-
ture, 323 K; mobile phase, 5 mM H2SO4; flow rate, 0.6 ml
min1; injection volume, 10 L. Using an external standard
method and calibration curves of commercially available
standard substrates, conversion (mol%) and product yield
(mol%) were calculated according to:

Fig. 1 Structural and textural properties of phosphated TiO2


nanoparticle: a) XRD pattern and b) N2 adsorptiondesorption isotherm.

This journal is The Royal Society of Chemistry 2016 Catal. Sci. Technol.
View Article Online

Paper Catalysis Science & Technology

thesis of phosphated TiO2 we reported previously41 can be


reduced by employing this simple one-pot route.

3.2. Activity tests


In a preliminary study, the catalytic activity of the as-
prepared phosphated TiO2 nanoparticle was evaluated for
glucose conversion into HMF in a waterMeTHF biphasic re-
action medium. Glucose conversion and HMF yield were
Published on 24 May 2016. Downloaded by University College London on 27/05/2016 10:49:57.

89.4% and 58.8%, respectively, as shown in Fig. 4. The cata-


lytic performance of phosphated TiO2 is ascribed to the pres-
ence of high surface acidity resulting from Lewis and
Brnsted acid sites, responsible for the tandem iso-
merisationdehydration reaction of glucose to HMF conver-
sion via fructose.41 Selective production of HMF was further
enhanced to a yield of 91% by modifying the biphasic me-
dium with N-methyl-pyrrolidone (NMP). NMP serves as an
Fig. 2 XPS analysis of phosphated TiO2 nanoparticle. a) Wide survey
aqueous phase modifier and its role is to a) suppress humin
scan; high resolution spectra of b) Ti 2p, c) P 2p, and d) O 1s.
formation and b) prevent rehydration of HMF.42 This is con-
firmed by the high concentration of HMF in the reactive
phase of the waterMeTHF system modified with NMP as
compared to that of the unmodified system (Fig. 4).
Furthermore, the product distribution shown in Fig. S2
confirmed that HMF undergoes rehydration to a greater ex-
tent in the unmodified reaction system, producing signifi-
cantly higher concentrations of levulinic acid and formic acid
as compared to the NMP modified system. Although, other
products identified were in minor concentrations, we de-
duced that the product balance is mainly humins, which was
remarkably high in the unmodified system and justified by
the colour change of the spent catalyst from white to intense
dark brown. Thus, the overall effectiveness of this biphasic
dehydration process towards selective HMF production is a
result of the cooperative effect of the high catalytic activity of
phosphated TiO2 and enhanced stability of HMF in the aque-
ous phase of the waterMeTHF/NMP system as well as good
partitioning of the formed HMF into the organic phase.

Fig. 3 Microscopic analysis of phosphated TiO2 showing micrographs


obtained by a) FE-SEM and b) HR-TEM.

estimated Ti and P content determined by ICP is 51.5 and Fig. 4 Glucose conversion to HMF in unmodified and modified water
MeTHF biphasic system. Reaction conditions: 10 g glucose, 1.25 g
3.49, respectively. Based on these data, we can conclude that
catalyst, 448 K, 80 min, 100 mL solvent volume (waterorganic solvent
the as-prepared material is an anatase polymorph with suc- 3 : 7 v/v). MeTHF : NMP 6 : 1 v/v for the modified reaction system.
cessful incorporation of phosphorus in the titania frame- Glucose conversion ( ) and HMF yield in organic phase () and
work. Therefore, the work-up procedure involved in the syn- aqueous phase ( ).

Catal. Sci. Technol. This journal is The Royal Society of Chemistry 2016
View Article Online

Catalysis Science & Technology Paper

The prospect of phosphated TiO2 as an effective bifunc- Table 2 Influence of cellulose depolymerization on HMF productiona
tional solid acid for the replacement of homogeneous catalyst
in the synthesis of HMF was explored. Under comparable re- Conversion Yield (%)
Entry Substrate (%) Glucose HMF
action conditions, we tested the catalytic dehydration of glu-
cose with homogeneous metal salts in the waterMTHF/NMP 1 Cellulose (MCC) 6.08 0.09 3.92
2 Ball-milled cellulose (BMC) 28.8 0.88 20.8
biphasic system and the results are presented in Table 1. 3 Mechanocatalytic depolymerized 99.8 0.16 74.7
AlCl3 exhibited the best catalytic performance, giving 72% cellulose (MDC)
yield of HMF at 45 min. Variation in the catalytic activity of a
Reaction conditions: 10 g substrate, 1.25 g catalyst, 453 K, 80 min,
the metal chlorides may be understood based on their Lewis 100 mL solvent volume (waterMeTHF/NMP 3 : 7 v/v).
Published on 24 May 2016. Downloaded by University College London on 27/05/2016 10:49:57.

acid softness and ionic radius as already described by Pagn-


Torres et al.26 Aluminium, being a very hard Lewis acid and
possessing a small ionic radius, has the strongest interaction
with glucose and hence, superior catalytic conversion to sion. This observation could be explained based on complete
HMF. The apparent higher yield of HMF arising from the ad- dissolution of cellulose in aqueous solution as a result of
dition of 0.1 M HCl to AlCl3 further confirms that the pres- decrystallization (Fig. S3) but most importantly, depolymeri-
ence of both Lewis and Brnsted acid sites play a crucial role zation to oligomers whose degree of polymerization (DP) is
in this reaction. According to the activity result, AlCl3/HCl less than 10.44
catalyst was able to convert glucose to 87% yield of HMF, Our result also indicates that reaction rate is dependent
which is comparable in magnitude to that obtained with the not only on the degree of cellulose crystallinity but strongly
phosphated TiO2 catalyst. on cleavage of the -1,4-glycosidic bond, which is considered
Next, we examined the reactivity of cellulose, a cheap and to be rate determining.48 Noteworthy to mention is the fact
readily available glucose polymer. Although conversion of cel- that the presence of residual acid in the acidulated cellulose
lulose is a challenge because of its chemical inertness and may promote HMF degradation. Shrotri et al.44 reported that
structural rigidity, structural deconstruction of cellulose is the presence of residual acid was responsible for the reduced
reported to improve its reactivity.46 As a control, conversion yield of sorbitol due to anhydrosorbitan formation as a result
of raw cellulose without any treatment was carried out and as of acid dehydration. To investigate this, an aqueous solution
expected, poor conversion and low yield of HMF were ob- of cello-oligomer was neutralized with barium hydroxide
served (Table 2, entry 1). For this reason, cellulose was (BaOH)2) prior to reaction. As shown in Fig. 5, HMF yield im-
subjected to ball milling prior to reaction. Conversion of cel- proved to 80.5%. Furthermore, when substrate concentration
lulose increased from 6.08% (raw cellulose; MCC) to 28.8% was changed from 10% to 5% loading, which corresponds to
(ball-milled cellulose; BMC). This may be explained as a re- pH variation from 1.27 to 1.65, HMF yield rose from 74.7% to
sult of decrystallization of MCC to the amorphous state (Fig. 83.5%.
S3), thus possessing fewer hydrogen bonds per repeating This result suggests that reducing the available residual
monomer.46,47 Consequently, MCC became sparingly soluble acid by lowering the substrate concentration can also mini-
in water after milling, thereby enhancing substratecatalyst mize HMF degradation. Further reduction of substrate con-
contact. Nevertheless, the solubility of milled cellulose in wa- centration to 1% (pH value of 2.25) resulted in a HMF yield
ter is low and as a result, milling alone was insufficient to of 93.5% and after neutralization, a similar yield (92.3%) of
achieve excellent cellulose reactivity as the yield of HMF does HMF was produced. Therefore, it seems that the most effec-
not exceed 20% (Table 2, entry 2). By acidulating cellulose be- tive way to inhibit HMF degradation is through neutraliza-
fore milling (mechanocatalytic depolymerized cellulose; tion, especially at high initial substrate concentration,
MDC) and then reacting the resultant substrate, 74.7% HMF
(Table 2, entry 3) was produced at a near-complete conver-

Table 1 Glucose conversion into HMF over homogeneous metal saltsa

Conversion Yield (%)


Entry Substrate (%) HMF Fructose AHG
1 AlCl36H2O 99.1 72.1 5.51 12.8
2 SnCl45H2O 85.6 51.0 23.5 5.48
3 CuCl22H2O 83.7 43.2 31.8 4.41
4 ZnCl2 78.6 28.2 34.6 3.59
5 AlCl36H2Ob 99.9 86.9 0.63 3.83
6 Phosphated TiO2c 97.9 90.7 0.37 0.96
a Fig. 5 Effect of residual acid on the conversion of cello-oligomers
Reaction conditions: 10 g glucose, 10 mM metal chloride, 448 K, 45
min, 100 mL solvent volume (waterMeTHF/NMP 3 : 7 v/v). b 0.1 M into HMF. Reaction conditions: 10 g substrate, 1.25 g catalyst, 448 K,
HCl co-catalyst. c 1.25 g cat. wt, 80 min reaction time. AHG, 80 min, 100 mL solvent volume (waterMeTHF/NMP 3 : 7 v/v). before
levoglucosan. neutralization, after neutralization.

This journal is The Royal Society of Chemistry 2016 Catal. Sci. Technol.
View Article Online

Paper Catalysis Science & Technology

considering that superior HMF yield was obtained after neu-


tralization at 5% (88.7% vs. 83.5%) and 10% (80.5% vs.
74.7%) substrate concentrations, respectively. More so,
irrespective of residual acid, substrate concentration is an-
other factor observed to affect the yield of HMF, which is in
agreement with previous studies.49 A steady loss in yield of
HMF as substrate concentration increased may be attributed
to a higher rate of cross-polymerization of HMF with reactive
intermediate compounds, thus forming undesired by-prod-
Published on 24 May 2016. Downloaded by University College London on 27/05/2016 10:49:57.

ucts. For example, in Fig. 5, at 1% substrate concentration


93.5% yield of HMF declined through 83.5% until it reached
74.7% when substrate concentration rose to 5% and 10%.
As we have shown that the phosphated TiO2 catalyst in Fig. 6 Cellulosic biomass conversion to furans on phosphated TiO2
combination with the waterMeTHF/NMP biphasic system catalyst in a waterMeTHF/NMP biphasic system. Reaction conditions:
was effective for the conversion of cellulose into a high yield 7.5 g substrate, 1.25 g catalyst, 453 K, 80 min, 100 mL solvent volume
(waterMeTHF/NMP 3 : 7 v/v). HMF, furfural.
of HMF, our strategy was then to seek a more economical
process for practical implementation by utilizing real bio-
mass as it serves as a naturally occurring renewable carbohy-
drate source. Hence, non-edible and inexpensive agricultural shown in Fig. 6. Chemical delignification of biomass appears
residues such as sugarcane bagasse and rice husk were con- to have a negligible influence in terms of total yield of furans
sidered in the present study. Prior to reaction, the biomass based on carbohydrate content, as results comparable to that
sources were washed with hot water, oven-dried at 323 K and of untreated biomass were obtained. This suggests that lignin
then milled in a Culatti Micro Hammer mill with 1.5 mm did not interfere substantially in the conversion of carbohy-
screen size. From the method established by Sluiter et al.,50 drates. Nonetheless, we anticipated that fractionating lignin
the carbohydrate component of bagasse is determined as will facilitate the ease of biomass deconstruction as lignin
41.5% cellulose and 24.1% xylan, whereas that of rice husk is content is reported to contribute to biomass recalcitrance
35.2% cellulose and 18.7% xylan. An integrated conversion of arising from the lignincarbohydrate interactions.54 Neverthe-
the biomass feedstocks to HMF and furfural, respectively, less, the composition of HMF in the furan mix produced after
was realized as shown in Fig. 6. For example, when bagasse biomass treatment is much higher, which evidently supports
was subjected to the transformation reaction, yields for HMF fractionation of the biomass substrates to give cellulose-rich
and furfural were 49.1% and 27.5%, respectively. We also pulp. Consequently, a gram of HMF produced per gram of
found that 51.9% and 23.1% yields for HMF and furfural, re- biomass consumed was effectively enhanced almost two-fold
spectively, could be produced from rice husk. as shown in Fig. S5. Therefore, a possible approach to sus-
To enhance the attractiveness of bio-refinery processes, it tainable HMF production is to adapt this catalytic process to
is important to fractionate biomass into its major compo- other agro-industrial residues as alternative feedstocks. As
nents, cellulose, hemicellulose and lignin, for effective valori- such, integration of this process to the existing agro-allied in-
zation into a variety of chemical compounds. This was dustry can potentially generate extra returns on investment
achieved by a two-stage fractionation process using alkali as well as simultaneously abate environmental concern and
pulping and then the OrganoCat process. Alkali treatment af- cost associated with waste disposal.
fords partial removal of lignin, xylan and other impurities
like pectin and wax.51,52 Another advantage of this step is the
weakening of the lignincarbohydrate bond linkage by esteri- 3.3. Catalyst recyclability
fication and swelling of the biomass fibres.51,53 Subsequently, A recycling test was performed with the catalyst to assess its
the alkali-treated biomass is subjected to the OrganoCat pro- reusability efficiency for the cello-oligomer to HMF reaction.
cess wherein xylan is hydrolysed into xylose and as a result, At the end of the reaction, the spent catalyst was recovered
lignin disentangles and gets separated into the organic phase from the reaction mixture by decantation, washed with ace-
to obtain a solid pulp rich in cellulose. As expected, cellulose tone, filtered and then dried at 373 K overnight prior to sub-
continuously increased upon chemical treatment. XRD data sequent reaction cycles. Fig. 7 shows that the HMF yield
of the biomass pulps (Fig. S4) show increased peak intensity monotonically declines from 94% in the fresh sample to
of cellulose after each stage of treatment, suggesting the re- about 80% by the end of the fourth reaction cycle. Partial loss
moval of xylan and lignin that are both of amorphous nature. of catalytic activity can be ascribed mainly to the active site
The recovered carbohydrate content of bagasse and husk af- blockage by humins arising from the discoloration of the cat-
ter the alkaliOrganoCat treatment is 90.96% (84.4% cellu- alyst from white to brown. In our previous study,41 a similar
lose, 6.56% xylan) and 83.94% (78% cellulose, 5.94% xylan), phenomenon was reported and the activity of the catalyst
respectively. The catalytic activity result for the conversion of could be restored by thermal treatment. Therefore, another
the resultant pulp after the alkaliorganocat treatment is reaction cycle was conducted (fifth reaction cycle) after the

Catal. Sci. Technol. This journal is The Royal Society of Chemistry 2016
View Article Online

Catalysis Science & Technology Paper

Scheme 1 Proposed reaction scheme of cellulose conversion to HMF


Published on 24 May 2016. Downloaded by University College London on 27/05/2016 10:49:57.

catalyzed by phosphated TiO2.

Fig. 7 Recyclability of phosphated TiO2 catalyst in the cello-oligomer


to HMF reaction. Reaction conditions: 1 g substrate, 1.25 g catalyst,
453 K, 80 min, 100 mL solvent volume (waterMeTHF/NMP 3 : 7 v/v). condensation of glucose as a result of its reversion reaction
gives glucose oligomers. In addition, glucose isomerizes to
fructose and subsequent elimination of three molecules of
spent catalyst had been thermally treated under the same water from the fructose molecule would then readily occur to
condition as the calcination treatment during catalyst prepa- form HMF. Fructose conversion to HMF rapidly takes place
ration, i.e. 873 K, 4 h. For the thermally treated spent cata- and occurs much faster than glucose isomerization to fruc-
lyst, HMF yield comparable to that of fresh catalyst was tose, and thus a high yield of fructose was not obtained at
achieved, thus suggesting that a simple regenerative protocol any time during the course of the reaction. As levulinic acid,
can be implemented to regain initial catalytic activity. formic acid, levoglucosan, and acetic acid were all observed
in minor quantities, they were not considered as part of the
main reactions of the proposed reaction scheme.
3.4. Reaction pathway and kinetics of cellulose to HMF Understanding the kinetics of HMF formation is essential
reaction to realizing effective catalytic technologies for the conversion
The reaction pathway of cellulose to HMF in water is consid- of carbohydrates. To the best of our knowledge, the kinetics
ered to be initiated by hydrolysis due to protonation of the of HMF formation in a biphasic system over a solid acid cata-
-1,4-glycosidic bonding followed by its dissociation to glu- lyst has never been reported. To this end, we studied the ki-
cose. Conversion of cellulose as a function of reaction time netics of phosphated TiO2 catalyzed conversion of cello-
was depicted by HPLC analysis and shown in Fig. S6a. Moni- oligomers in a waterMeTHF biphasic system. Cello-oligomer
toring the evolution of products as the reaction proceeds, we (i.e. depolymerized cellulose) was chosen as a model to repre-
noticed peaks identified as glucose and cellobiose in the first sent cellulose because it is soluble and has better reactivity
few minutes of the reaction. Additionally, the HPLC chro- as earlier observed (Table 2). Based on our result, the main
matograph revealed extra peaks (1 and 3) which are products detected are glucose and HMF. Hence, cello-
unidentified but are likely to be glucose oligomers. This is oligomer hydrolysis to glucose and glucose dehydration to
because cleavage of the -1,4-glycosidic bond is generally ac- HMF is considered as a consecutive first-order reaction in ac-
cepted to produce water-soluble oligomers prior to glucose cordance with the widely accepted model reported by many
formation.55 The observance of peak 3 as well as cellobiose authors.5860 Rate constants for glucose isomerization to fruc-
during the glucose conversion reaction (Fig. S6c) suggests tose and the reversible reaction are both neglected. We as-
reversion of glucose due to self-condensation.56,57 Likewise, sumed that fructose is dehydrated much rapidly to HMF as it
peak 3 was also observed during cellobiose reaction, which is formed, and as a result only a trace concentration of fruc-
further supports the assumption that the unknown peak may tose is present in the product mix. Glucose oligomers are also
be a glucose oligomer. Additional peaks, 6 and 7, noticeable detected in trace amounts under this biphasic reaction condi-
in glucose and cellobiose but vaguely present in cellulose, tion; therefore, oligomerization of glucose is neglected. Like-
may be regarded as glucose decomposition products. This as- wise, other products such as levulinic acid, formic acid, and
sumption was based on the fact that when fructose was levoglucosan, did not exceed 2% yield; hence, they are not
reacted as shown in Fig. S6d, these peaks were not present. accounted for. Thus, we simplify the cello-oligomer conver-
However, no attempts were made to elucidate on either the sion reaction in our reaction system as:
pathway of glucose decomposition or that of subsequent
HMF degradation. Hence, a plausible reaction pathway for
the phosphated TiO2 catalyzed cellulose to HMF conversion in which k1 and k2 are reaction rate constants for the hydroly-
in a waterMeTHF biphasic system is shown in Scheme 1. sis and dehydration steps, respectively.
Hydrolysis of cellulose is assumed to proceed mainly with the To estimate the kinetic parameters, we employ a kinetic
formation of oligosaccharides. Further hydrolysis of the oli- model that consists of two reaction steps: i) cello-oligomer
gosaccharides eventually yields glucose. Meanwhile, self- hydrolytic step and ii) glucose dehydration step. As a starting

This journal is The Royal Society of Chemistry 2016 Catal. Sci. Technol.
View Article Online

Paper Catalysis Science & Technology

point, glucose dehydration experiments were carried out over


a reaction temperature range of 428448 K with an initial glu-
cose concentration of 10% (w/v) and sampling interval of 15
min from 0 to 60 min. The reaction rate equations of the
glucose-to-HMF step are:

(1)
Published on 24 May 2016. Downloaded by University College London on 27/05/2016 10:49:57.

(2)

Similar reaction conditions were used for the cello-oligomer


to HMF reaction, and the following are the derived reaction
rate equations:

(3)

(4)

(5)
Fig. 8 a) Arrhenius plot of cello-oligomer conversion to HMF: cello-
oligomer hydrolysis to glucose (blue), glucose dehydration to HMF (ol-
where Cg, CH and Cc are the concentrations of glucose, HMF ive green). b) Parity plot of experimental data and model prediction;
and cello-oligomers. glucose yield, HMF yield. Reaction conditions: 428448 K, 10% (w/v)
initial substrate concentration, 8 : 1 substrate/cat wt ratio, 100 mL sol-
Experimental rate data of cello-oligomer conversion and
vent (waterMeTHF/NMP 3 : 7 v/v).
yields of both glucose and HMF employed in the formulation
of the reaction model are depicted in Fig. S7. The resulting
set of ordinary differential equations (ODEs) was simulta-
neously solved by a MATLAB program. A least-squares algo- fore, the estimated kinetic data for cello-oligomer conversion
rithm was adopted to minimize the error between experimen- to HMF in this biphasic reaction system provides a reason-
tal and predicted data and then subsequent evaluation of able estimation that is in fair agreement with those reported
reaction rate constants through curve fitting. Activation ener- in the literature.
gies were calculated using the Arrhenius equation given as: For instance, Girisuta et al.62 reported that cellulose and
glucose undergo hydrolysis and degradation reaction, respec-
(6) tively, with similar activation energies of approximately 152
kJ mol1 from a detailed kinetic study of acid-catalyzed hydro-
lysis reaction of cellulose in the presence of sulphuric acid.
The Arrhenius rate data plot is shown in Fig. 8a. As a mea- Jiang et al.63 obtained activation energies of 114 and 95 kJ
sure of good approximation of the mathematical model to fit mol1 corresponding to cellulose hydrolysis and glucose
experimental data, the R-squared value for both sets of reac-
tion steps is greater than 0.9. The result of the kinetic study
is presented in Table 3 for which activation energies of 110.9 Table 3 Estimated kinetic parameters for cello-oligomer and glucose
kJ mol1 and 122.4 kJ mol1 were calculated for both the de- conversion to HMF in a waterMeTHF/NMP biphasic system
hydration and the hydrolysis steps, respectively. The order of
Temperature K
magnitude of these energies is similar and this was also
Rate constant (min1) 428 438 448
reported by Dee et al.61 According to the authors, these en-
ergy barriers are insensitive to temperature and their differ- k1 0.0206 0.0352 0.0961
k2 0.0117 0.0225 0.0471
ence is within 711 kJ mol1. Furthermore, the simplicity of
Activation energy (kJ mol1)
the reaction model remarkably predicts the experimentally E1 122.4
observed data as shown by the parity plot of Fig. 8b. There- E2 110.9

Catal. Sci. Technol. This journal is The Royal Society of Chemistry 2016
View Article Online

Catalysis Science & Technology Paper

dehydration steps, respectively, using acidic ionic liquid cata- 7 T. Wang, M. W. Nolte and B. H. Shanks, Green Chem.,
lyst in 1-butyl-3-methylimidazolium chloride ([Bmim]Cl) sol- 2014, 16, 548572.
vent. Using cellobiose as a model compound for cellulose, 8 R. L. de Souza, H. Yu, F. Rataboul and N. Essayem,
Vanoye et al.60 determined the activation energies of cellobi- Challenges, 2012, 3, 212232.
ose hydrolysis and glucose dehydration in [C2mim]Cl IL as 9 Y. Zhang, J. Pan, Y. Shen, W. Shi, C. Liu and L. Yu, ACS
111 kJ mol1 and 102 kJ mol1, respectively. Sustainable Chem. Eng., 2015, 3, 871879.
10 Y. Zhang, H. Du, X. Qian and E. Y.-X. Chen, Energy Fuels,
4. Conclusions 2010, 24, 24102417.
11 Z. Zhang and Z. K. Zhao, Bioresour. Technol., 2010, 101,
We have described a simple and efficient process suited for
Published on 24 May 2016. Downloaded by University College London on 27/05/2016 10:49:57.

11111114.
the conversion of biomass-derived carbohydrates into HMF. 12 A. M. da Costa Lopes and R. Bogel-ukasik, ChemSusChem,
The preparation of phosphated TiO2 was achieved by a one- 2015, 8, 947965.
pot synthesis with anatase as the main phase of TiO2 but 13 J. B. Binder and R. T. Raines, J. Am. Chem. Soc., 2009, 131,
modified by the formation of a TiOP bonding. At 10% (w/v) 19791985.
substrate loading, HMF yields of 91% and 74.7% from glu- 14 Y. Su, H. M. Brown, X. Huang, X.-d. Zhou, J. E. Amonette
cose and solubilized cellulose, respectively, were achieved in and Z. C. Zhang, Appl. Catal., A, 2009, 361, 117122.
a NMP modified waterMeTHF biphasic medium. Reaction 15 B. Kim, J. Jeong, D. Lee, S. Kim, H.-J. Yoon, Y.-S. Lee and
kinetics for the formation of HMF was studied. By analyzing J. K. Cho, Green Chem., 2011, 13, 15031506.
experimental data, a simple kinetic model was developed to 16 P. Wang, H. Yu, S. Zhan and S. Wang, Bioresour. Technol.,
estimate kinetic parameters for the tandem hydrolysisdehy- 2011, 102, 41794183.
dration reaction. The activation energy of cello-oligomer hy- 17 W.-H. Peng, Y.-Y. Lee, C. Wu and K. C. W. Wu, J. Mater.
drolysis is 122.4 kJ mol1 and the energy required for the sub- Chem., 2012, 22, 2318123185.
sequent dehydration of glucose to HMF is 110.9 kJ mol1. 18 I. J. Kuo, N. Suzuki, Y. Yamauchi and K. C. W. Wu, RSC Adv.,
Another desirable attribute of our process is the significant 2013, 3, 20282034.
conversion of real biomass to both HMF and furfural. A com- 19 W.-H. Hsu, Y.-Y. Lee, W.-H. Peng and K. C. W. Wu, Catal.
bined furan yield of about 76% (4952% HMF and 2328% Today, 2011, 174, 6569.
furfural) is achievable. Catalytic conversion of the cellulose 20 Y. Zhang, Y. Chen, Y. Shen, Y. Yan, J. Pan, W. Shi and L. Yu,
rich pulp after biomass fractionation led to an increased pro- ChemPlusChem, 2016, 81, 108118.
duction in HMF to about 72% (bagasse) and 65% (husk). 21 S. P. Teong, G. Yi and Y. Zhang, Green Chem., 2014, 16,
Overall, the method described herein serves as a means of 20152026.
finding a feasible economical and environmentally friendly 22 Y. Romn-Leshkov, J. N. Chheda and J. A. Dumesic, Science,
industrial process for the production of HMF. 2006, 312, 19331937.
23 Y. Yang, C.-w. Hu and M. M. Abu-Omar, Green Chem.,
Acknowledgements 2012, 14, 509513.
The authors gratefully acknowledge the financial support of 24 N. Shi, Q. Liu, Q. Zhang, T. Wang and L. Ma, Green Chem.,
the Nanomaterials Centre at The University of Queensland, 2013, 15, 19671974.
Australia. The authors also acknowledge the use of facilities 25 B. Saha and M. M. Abu-Omar, Green Chem., 2014, 16, 2438.
and the scientific and technical assistance of the Australian 26 Y. J. Pagan-Torres, T. Wang, J. M. R. Gallo, B. H. Shanks and
Microscopy & Microanalysis Research Facility at The Univer- J. A. Dumesic, ACS Catal., 2012, 2, 930934.
sity of Queensland. The first author sincerely appreciates the 27 T. Wang, Y. J. Pagn-Torres, E. J. Combs, J. A. Dumesic and
sponsorship from the International Postgraduate Research B. H. Shanks, Top. Catal., 2012, 55, 657662.
Scholarship (IPRS) and UQ Centennial Scholarship (UQ Cent). 28 T. Dallas Swift, H. Nguyen, A. Anderko, V. Nikolakis and
D. G. Vlachos, Green Chem., 2015, 17, 47254735.
Notes and references 29 R. Carrasquillo-Flores, M. Kldstrm, F. Schth, J. A.
Dumesic and R. Rinaldi, ACS Catal., 2013, 3, 993997.
1 G. W. Huber, S. Iborra and A. Corma, Chem. Rev., 2006, 106, 30 R. Otomo, T. Yokoi, J. N. Kondo and T. Tatsumi, Appl.
40444098. Catal., A, 2014, 470, 318326.
2 J. C. Serrano-Ruiz and J. A. Dumesic, Energy Environ. Sci., 31 E. Nikolla, Y. Romn-Leshkov, M. Moliner and M. E. Davis,
2011, 4, 8399. ACS Catal., 2011, 1, 408410.
3 P. Gallezot, Chem. Soc. Rev., 2012, 41, 15381558. 32 C. Fan, H. Guan, H. Zhang, J. Wang, S. Wang and X. Wang,
4 L. Faba, E. Daz and S. Ordez, Renewable Sustainable Biomass Bioenergy, 2011, 35, 26592665.
Energy Rev., 2015, 51, 273287. 33 Q. Zhao, L. Wang, S. Zhao, X. Wang and S. Wang, Fuel,
5 A. A. Rosatella, S. P. Simeonov, R. F. M. Frade and C. A. M. 2011, 90, 22892293.
Afonso, Green Chem., 2011, 13, 754793. 34 Y. Lu, Z. Sun and M. Huo, RSC Adv., 2015, 5, 3086930876.
6 S. Dutta, S. De and B. Saha, ChemPlusChem, 2012, 77, 35 S. Mondal, J. Mondal and A. Bhaumik, ChemCatChem,
259272. 2015, 7, 35703578.

This journal is The Royal Society of Chemistry 2016 Catal. Sci. Technol.
View Article Online

Paper Catalysis Science & Technology

36 V. V. Ordomsky, J. van der Schaaf, J. C. Schouten and T. A. 51 X. Zhao, F. Peng, K. Cheng and D. Liu, Enzyme Microb.
Nijhuis, ChemSusChem, 2013, 6, 16971707. Technol., 2009, 44, 1723.
37 M. I. Alam, S. De, B. Singh, B. Saha and M. M. Abu-Omar, 52 N. Johar, I. Ahmad and A. Dufresne, Ind. Crops Prod.,
Appl. Catal., A, 2014, 486, 4248. 2012, 37, 9399.
38 S. Dutta, S. De, A. K. Patra, M. Sasidharan, A. Bhaumik and 53 M. M. Ibrahim, W. K. El-Zawawy, Y. R. Abdel-Fattah, N. A.
B. Saha, Appl. Catal., A, 2011, 409410, 133139. Soliman and F. A. Agblevor, Carbohydr. Polym., 2011, 83,
39 A. Dutta, A. K. Patra, S. Dutta, B. Saha and A. Bhaumik, 720726.
J. Mater. Chem., 2012, 22, 1409414100. 54 M. E. Himmel, S.-Y. Ding, D. K. Johnson, W. S. Adney, M. R.
40 A. Dutta, D. Gupta, A. K. Patra, B. Saha and A. Bhaumik, Nimlos, J. W. Brady and T. D. Foust, Science, 2007, 315,
Published on 24 May 2016. Downloaded by University College London on 27/05/2016 10:49:57.

ChemSusChem, 2014, 7, 925933. 804807.


41 L. Atanda, S. Mukundan, A. Shrotri, Q. Ma and J. Beltramini, 55 S. Van de Vyver, J. Thomas, J. Geboers, S. Keyzer, M. Smet,
ChemCatChem, 2015, 7, 781790. W. Dehaen, P. A. Jacobs and B. F. Sels, Energy Environ. Sci.,
42 L. Atanda, A. Shrotri, S. Mukundan, Q. Ma, M. Konarova and 2011, 4, 36013610.
J. Beltramini, ChemSusChem, 2015, 8, 29072916. 56 Q. Xiang, Y. Y. Lee and R. W. Torget, in Proceedings of the
43 P. M. Grande, J. Viell, N. Theyssen, W. Marquardt, P. Twenty-Fifth Symposium on Biotechnology for Fuels and
Dominguez de Maria and W. Leitner, Green Chem., 2015, 17, Chemicals Held May 47, 2003, in Breckenridge, CO, ed. M.
35333539. Finkelstein, J. D. McMillan, B. H. Davison and B. Evans,
44 A. Shrotri, L. K. Lambert, A. Tanksale and J. Beltramini, Humana Press, Totowa, NJ, 2004, pp. 11271138.
Green Chem., 2013, 15, 27612768. 57 H. M. Pilath, M. R. Nimlos, A. Mittal, M. E. Himmel and
45 J. C. Yu, L. Zhang, Z. Zheng and J. Zhao, Chem. Mater., D. K. Johnson, J. Agric. Food Chem., 2010, 58, 61316140.
2003, 15, 22802286. 58 J. F. Saeman, Ind. Eng. Chem., 1945, 37, 4352.
46 P. L. Dhepe and A. Fukuoka, ChemSusChem, 2008, 1, 969975. 59 N. S. Mosier, A. Sarikaya, C. M. Ladisch and M. R. Ladisch,
47 K. Mazeau and L. Heux, J. Phys. Chem. B, 2003, 107, Biotechnol. Prog., 2001, 17, 474480.
23942403. 60 L. Vanoye, M. Fanselow, J. D. Holbrey, M. P. Atkins and K. R.
48 A. Cabiac, E. Guillon, F. Chambon, C. Pinel, F. Rataboul and Seddon, Green Chem., 2009, 11, 390396.
N. Essayem, Appl. Catal., A, 2011, 402, 110. 61 S. J. Dee and A. T. Bell, ChemSusChem, 2011, 4, 11661173.
49 B. Kuster, Starch-Strke, 1990, vol. 42, pp. 314321. 62 B. Girisuta, L. Janssen and H. Heeres, Ind. Eng. Chem. Res.,
50 A. Sluiter, B. Hames, R. Ruiz, C. Scarlata, J. Sluiter, D. 2007, 46, 16961708.
Templeton and D. Crocker, Report N. TP-510-42618, 2011, 63 F. Jiang, Q. Zhu, D. Ma, X. Liu and X. Han, J. Mol. Catal. A:
p. 17. Chem., 2011, 334, 812.

Catal. Sci. Technol. This journal is The Royal Society of Chemistry 2016

Das könnte Ihnen auch gefallen