Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

catalysts

Article
Conversion of Cellulose to Lactic Acid by Using
ZrO2–Al2O3 Catalysts
Panya Wattanapaphawong 1,2 , Osamu Sato 1 , Koichi Sato 1 , Naoki Mimura 1 ,
Prasert Reubroycharoen 2,3 and Aritomo Yamaguchi 1, *
1 Research Institute for Chemical Process Technology, National Institute of Advanced Industrial Science and
Technology (AIST), 4-2-1 Nigatake, Miyagino, Sendai 983-8551, Japan; yoyo_yok7@hotmail.com (P.W.);
o.satou@aist.go.jp (O.S.); koichi.sato@aist.go.jp (K.S.); n.mimura@aist.go.jp (N.M.)
2 Department of Chemical Technology, Faculty of Science, Chulalongkorn University, Pathumwan,
Bangkok 10330, Thailand; prasert.r@chula.ac.th
3 Center of Excellence on Petrochemical and Materials Technology,
Chulalongkorn University Research Building, Bangkok 10330, Thailand
* Correspondence: a.yamaguchi@aist.go.jp; Fax: +81-22-237-5226

Received: 26 June 2017; Accepted: 19 July 2017; Published: 21 July 2017

Abstract: Lactic acid has a wide range of applications in many industries, both as an ingredient and
as an intermediate. Here, we investigated the catalytic conversion of cellulose to lactic acid by using
heterogeneous mixed-oxide catalysts containing ZrO2 . Although pure ZrO2 has catalytic activity for
the conversion of cellulose to lactic acid, the yield of lactic acid obtained is not satisfactory. In contrast,
a series of ZrO2 –Al2 O3 catalysts containing various percentages of ZrO2 provided higher yields of
lactic acid. The ZrO2 –Al2 O3 catalysts had more Lewis acid sites and far fewer base sites than ZrO2 .
This suggests that the Lewis acid sites on ZrO2 –Al2 O3 catalysts are more important than the base
sites for the conversion of cellulose to lactic acid.

Keywords: cellulose; lactic acid; ZrO2 –Al2 O3 ; zirconium oxide; aluminum oxide; biomass

1. Introduction
Biomass is a renewable organic resource that contains components that are essential for the
production of many important chemicals and fuels [1–3]. Inedible lignocellulosic biomass is plant
biomass that is mainly composed of cellulose, hemicellulose, and lignin. Currently, cellulose is the
most promising component for biomass utilization because cellulose is the major component of
lignocellulosic biomass [4–7].
Lactic acid is an organic compound that is used in large quantities as the food, cosmetic,
pharmaceutical, and chemical productions [8,9]. Moreover, lactic acid is an important intermediate
for conversion to other products such as propylene glycol, acrylic acid, and polylactic acid [8,10].
Conventionally, lactic acid is produced via the fermentation of sugars from starch [11]; however, the rate
of production using this method is low, and the pH of the solution during fermentation must be kept
in the range 5–7 to provide an environment suitable for lactic acid-fermenting bacteria. Thus, catalytic
conversion of inedible cellulose is an attractive alternative to fermentation for the production of
lactic acid.
Recently, several homogeneous catalysts (e.g., PbCl2 and ErCl3 ) have been investigated for their
suitability for the conversion of cellulose to lactic acid [12,13]. These catalysts provide a high yield of
lactic acid; however, their use in industries remains unrealized because of difficult separation of the
products from the catalyst as well as the poor stability and recyclability of the catalysts. Therefore,
heterogeneous catalysts for the conversion of cellulose to lactic acid are desired because products can
be easily separated from this type of catalyst. Several heterogeneous catalysts have been reported to

Catalysts 2017, 7, 221; doi:10.3390/catal7070221 www.mdpi.com/journal/catalysts


Catalysts 2017, 7, 221 2 of 10

have catalytic activity for the conversion of cellulose to lactic acid, such as LaCoO3 , NbF5 –AlF3 , and
AlW, which provided
Catalysts 2017, 7, 221yields of lactic acid of 24%, 27%, and 28%, respectively [14–16]. However, 2 of 10 these
solid catalysts have a disadvantage that the metal species leach into solution during the reaction—2.4%
of the Cohave
andbeen
1.5%reported
of the La to from
have catalytic
LaCoO3activity
[14] andfor 1.5%
the conversion
of the W of cellulose
from AlW to lactic
[16] acid, such
leached into assolution.
LaCoO3, NbF5–AlF3, and AlW, which provided yields of lactic acid of 24%, 27%, and 28%,
Additionally, these heterogeneous catalysts are relatively expensive, making their industrial use
respectively [14–16]. However, these solid catalysts have a disadvantage that the metal species leach
difficult.into
Previously, we reported
solution during a simple of
the reaction—2.4% ZrO catalyst
the2 Co forofthe
and 1.5% theconversion
La from LaCoO of cellulose to lactic acid
3 [14] and 1.5% of

(yield, 21.2%) that did not leach metal species into the reaction solution
the W from AlW [16] leached into solution. Additionally, these heterogeneous catalysts are [17]. In addition, the ZrO2
catalyst was stable in hot water. However, the yield of lactic acid obtained with this catalyst was
relatively expensive, making their industrial use difficult. Previously, we reported a simple ZrO 2 less
catalyst for the conversion of cellulose to lactic acid (yield, 21.2%) that did
than those reported for the heterogeneous catalysts. Thus, the aim of the present study was to examine not leach metal species
into the reaction solution [17]. In addition, the ZrO2 catalyst was stable in hot water. However, the
how to increase the yield of lactic acid obtainable with ZrO2 -based catalysts. Dispersion of active
yield of lactic acid obtained with this catalyst was less than those reported for the heterogeneous
species on supports such as metal oxides is a common means of increasing the catalytic activity of
catalysts. Thus, the aim of the present study was to examine how to increase the yield of lactic acid
heterogeneous
obtainable catalysts.
with ZrO Therefore, here we
2‐based catalysts. compared
Dispersion the catalytic
of active species onactivity
supportsofsucha series of mixed-oxide
as metal oxides
catalystsiscontaining various amounts of
a common means of increasing the catalytic ZrO 2 (5%, 10%, or 20%) on an Al O
activity of heterogeneous2catalysts.
3 supportTherefore,that
with hereof pure
we compared the catalytic activity of a series of mixed‐oxide catalysts containing
ZrO2 . We found that the 10%ZrO2 –Al2 O3 catalyst showed activity for the production of lactic acid from various amounts
celluloseofwith
ZrO2a(5%, 10%, or
product 20%)ofon25.3%,
yield an Al2O 3 support with that of pure ZrO2. We found that the 10%ZrO2–
which was cheaper than the reported heterogeneous catalysts.
Al2O3 catalyst showed activity for the production of lactic acid from cellulose with a product yield of
2. Results andwhich
25.3%, was cheaper than the reported heterogeneous catalysts.
Discussion
2. Results and Discussion
2.1. Effect of Ball-Milling Treatment on Cellulose Conversion
2.1.paper,
In this Effect ofpulverization
Ball‐Milling Treatment on Cellulose
of cellulose Conversion has been applied for the increase of cellulose
by ball-milling
conversion [18–20]. The conversion
In this paper, pulverization of of
cellulose
cellulosewas only 25% in
by ball‐milling hasthe caseapplied
been of un-milled cellulose
for the increase of(473 K,
cellulose
6 h, without conversion
catalysts) [18–20]. Thetoconversion
and increased 86% by the of ball-milling
cellulose was treatment
only 25% in ofthe case of un‐milled
cellulose. Figure 1 shows
cellulose (473 K, 6 h, without catalysts) and increased to 86% by the ball‐milling treatment of
XRD patterns of un-milled and milled cellulose. The cellulose crystallinity was decreased by the
cellulose. Figure 1 shows XRD patterns of un‐milled and milled cellulose. The cellulose crystallinity
ball-milling treatment as revealed by the peak broadening at 2θ = 22.6 degree (crystalline plane 002)
was decreased by the ball‐milling treatment as revealed by the peak broadening at 2θ = 22.6 degree
in XRD patterns (Figure
(crystalline 1) [19,21],
plane 002) in XRDindicating that the
patterns (Figure cellulose
1) [19,21], conversion
indicating increased
that the with decreasing
cellulose conversion
celluloseincreased
crystallinity. In the reaction
with decreasing cellulosepathway, cellulose
crystallinity. In thewas firstly
reaction hydrolyzed
pathway, to was
cellulose soluble glucose;
firstly
hydrolyzed
thus, cellulose withtohigh
soluble glucose; thus,
crystallinity was cellulose with high
hydrolyzed crystallinity
very slowly atwas473hydrolyzed
K from the very slowly
result ofatthe low
473 of
conversion K from the result
un-milled of the low conversion of un‐milled cellulose.
cellulose.

Figure 1. XRD patterns of (a) un‐milled cellulose and (b) milled cellulose.
Figure 1. XRD patterns of (a) un-milled cellulose and (b) milled cellulose.
2.2. Catalytic Conversion of Cellulose to Lactic Acid by Using ZrO2–Al2O3
2.2. Catalytic Conversion of Cellulose to Lactic Acid by Using ZrO2 –Al2 O3
Previously, we investigated the catalytic conversion of cellulose to lactic acid by using various
transitionwe
Previously, metal oxides (i.e., ZrO
investigated the2,catalytic
Al2O3, TiO2, Fe3O4, V2Oof
conversion 5, CeO 2, Y2O3, to
cellulose Tmlactic
2O3, HfO2, Ga
acid by2Ousing
3, MgO,various
La 2O3, Nb2O5, and Ta2O5) [17]. Of these catalysts, ZrO2 had the highest catalytic activity (lactic acid
transition metal oxides (i.e., ZrO2 , Al2 O3 , TiO2 , Fe3 O4 , V2 O5 , CeO2 , Y2 O3 , Tm2 O3 , HfO2 , Ga2 O3 ,
yield, 21.2%; reaction temperature, 473 K). Therefore, in the present study, we examined the use of
MgO, La2 O3 , Nb2 O5 , and Ta2 O5 ) [17]. Of these catalysts, ZrO2 had the highest catalytic activity
(lactic acid yield, 21.2%; reaction temperature, 473 K). Therefore, in the present study, we examined
the use of an Al2 O3 support to increase the catalytic activity of ZrO2 for the conversion of cellulose
to lactic acid. The yields of lactic acid obtained with ZrO2 –Al2 O3 mixed-oxide catalysts containing
Catalysts
Catalysts 2017, 7, 2017,
221 7, 221 3 of 10 3 of 10

an Al2O3 support to increase the catalytic activity of ZrO2 for the conversion of cellulose to lactic
Catalysts 2017, 7, 221 3 of 10
acid. The yields of lactic acid obtained with ZrO2–Al2O3 mixed‐oxide catalysts containing different
different percentages of ZrO2 at 473 K were higher than those obtained by using the individual metal
percentages of ZrO2 at 473 K were higher than those obtained by using the individual metal oxides,
an Al2O3 support
oxides,indicating
indicating that theto increase the catalytic
dispersion activity
of active ZrOof ZrO 2 for the
species on conversion
the Al2 O3ofsupport
cellulose increased
to lactic the
that the dispersion of active ZrO2 species 2on the Al2O3 support increased the catalytic
catalytic acid. The
activity yields
of ZrO of lactic acid obtained with ZrO 2–Al2O3 mixed‐oxide catalysts containing different
activity of ZrO 2 (Figure
2 (Figure 2). The2). The maximum
maximum yield
yield of lactic acidof lacticwas
(25.3%) acidobtained
(25.3%)when
was10%ZrO
obtained 2–
when
percentages of ZrO2 at 473 K were higher than those obtained by using the individual metal oxides,
10%ZrO Al2 2–Al
O3 wasO was
used used
as the as the
catalyst. catalyst.
The yield The
of yield
lactic acidof lactic
increased acid increased
slightly as the slightly
amount
2 3 that the dispersion of active ZrO2 species on the Al2O3 support increased the catalytic of as
ZrO the
2 on amount
indicating
of ZrO2theon Althe
2O3 Al
activity
support
of ZrO
was increased increased
2 O32 support
(Figure 2).was
from 5% to from
The maximum yield
10%; however,
of 5% toacid
lactic
the yield
10%; decreased
however,
(25.3%) the as
was obtained
the amount
yield
when decreased of
10%ZrO2– as the
ZrO
amount of 2 was increased to 20%, indicating that the yield of lactic acid may decrease further as the
Al2ZrO 2 was
O3 was used increased to 20%,
as the catalyst. Theindicating that
yield of lactic theincreased
acid yield ofslightly
lactic acid
as themay decrease
amount of ZrOfurther
2 on as
percentage of ZrO2 on the support increases.
the percentage
the Al2O of3 support
ZrO2 on theincreased
was supportfromincreases.
5% to 10%; however, the yield decreased as the amount of
ZrO2 was increased to 20%, indicating that the yield of lactic acid may decrease further as the
percentage of ZrO2 on the support increases.

Figure 2. Product yields obtained by using Al2O3, ZrO2, or ZrO2–Al2O3 mixed‐oxide catalysts for the
Figure 2. Product yields obtained by using Al2 Oconditions:
conversion of cellulose to lactic acid. Reaction 3 , ZrO2 , or ZrO2 –Alcellulose,
ball‐milled 2 O3 mixed-oxide catalysts for the
0.5 g; catalyst, 1.0 g;
conversion
water, 50 g; reaction temperature, 473 K; reaction time, 6 h. HMF, 5‐hydroxymethylfurfural. catalyst, 1.0 g;
of cellulose to lactic acid. Reaction conditions: ball-milled cellulose, 0.5 g;
Figure 2. Product yields obtained by using Al2O3, ZrO2, or ZrO2–Al2O3 mixed‐oxide catalysts for the
water, 50 g; reaction temperature, 473 K; reaction time, 6 h. HMF, 5-hydroxymethylfurfural.
conversion of cellulose to lactic acid. Reaction conditions: ball‐milled cellulose, 0.5 g; catalyst, 1.0 g;
2.3. Reaction Conditions
water, 50 g; reaction temperature, 473 K; reaction time, 6 h. HMF, 5‐hydroxymethylfurfural.
2.3. ReactionNext,
Conditions
we optimized the reaction conditions to obtain the highest possible yields of lactic acid
2.3. 10%ZrO
with Reaction2Conditions
–Al2O3 used as the catalyst. Figure 3 shows the product yields obtained as a function
Next, we optimized
of reaction time. The the
yieldreaction
of lactic conditions
acid increasedto for
obtain
up tothe
6 h,highest possible
after which yieldsdue
it decreased of lactic
to acid
Next, we optimized the reaction conditions to obtain the highest possible yields of lactic acid
with 10%ZrO –Al
decomposition
2 O
2 or used as the catalyst. Figure
3 polymerization of lactic acid. 3 shows the product yields obtained as a function
with 10%ZrO2–Al2O3 used as the catalyst. Figure 3 shows the product yields obtained as a function
of reaction time. The yield of lactic acid increased for up to 6 h, after which it decreased due to
of reaction time. The yield of lactic acid increased for up to 6 h, after which it decreased due to
decomposition or polymerization
decomposition or polymerization of lactic acid.
of lactic acid.

Figure 3. Product yields obtained when 10%ZrO2 –Al2 O3 was used as the catalyst as a function of
reaction time. Reaction conditions: ball-milled cellulose, 0.5 g; catalyst, 1.0 g; water, 50 g; reaction
temperature, 473 K. HMF, 5-hydroxymethylfurfural.
Catalysts 2017, 7, 221 4 of 10

Figure 3. Product yields obtained when 10%ZrO2–Al2O3 was used as the catalyst as a function of
reaction
Catalysts time. Reaction conditions: ball‐milled cellulose, 0.5 g; catalyst, 1.0 g; water, 50 g; reaction
2017, 7, 221 4 of 10
temperature, 473 K. HMF, 5‐hydroxymethylfurfural.

Figure
Figure4 shows
4 shows the the
product yieldsyields
product obtained as a function
obtained of reaction
as a function oftemperature. The maximum
reaction temperature. The
yield of lacticyield
maximum acidof(25.3%)
lactic was
acidobtained at 473
(25.3%) was K. At theatlowest
obtained 473 K.temperature examined
At the lowest (i.e., 453
temperature K), the
examined
cellulose
(i.e., 453conversion reactionconversion
K), the cellulose proceeded very slowly.
reaction At the highest
proceeded temperature
very slowly. At theexamined (i.e., 493 K),
highest temperature
examined
the (i.e., 493
yield of lactic acid K),
wasthe
lessyield of lactic
than that at the acid was temperature
optimum less than that at thethe
because optimum
lactic acidtemperature
was likely
because the lactic acid was likely decomposed at this temperature.
decomposed at this temperature.

Figure 4. Product yields obtained when 10%ZrO2–Al2O3 was used as the catalyst as a function of
Figure 4. Product yields obtained when 10%ZrO2 –Al2 O3 was used as the catalyst as a function of
reaction temperature. Reaction conditions: ball‐milled cellulose, 0.5 g; catalyst, 1.0 g; water, 50 g;
reaction temperature. Reaction conditions: ball-milled cellulose, 0.5 g; catalyst, 1.0 g; water, 50 g;
reaction time, 6 h. HMF, 5‐hydroxymethylfurfural.
reaction time, 6 h. HMF, 5-hydroxymethylfurfural.

Finally, we investigated the reusability of the 10%ZrO2–Al2O3 catalyst for repeated conversion
Finally, we
of cellulose toinvestigated
lactic acid. the reusability
After of the 10%ZrO
the reaction with the2 –Al 2 O3 catalyst
optimized for repeated
conditions (473conversion
K, 6 h), theof
cellulose to lactic acid. After the reaction with the optimized conditions (473
recovered solid, which contained the used 10%ZrO2–Al2O3 catalyst and any unreacted cellulose, K, 6 h), the recovered
solid,
was which
heatedcontained
at 673 K thefor
used
1510%ZrO
h under 2 –Alan
2 O3air
catalyst and anyto
atmosphere unreacted
remove cellulose, was heated
carbon‐based at
material
673 K for 15 h under an air atmosphere to remove carbon-based material accumulated
accumulated on the surface of the catalyst. The yield of lactic acid obtained by using the reused on the surface of
the catalyst.
10%ZrO The yield of lactic acid obtained by using the reused 10%ZrO2 –Al2the
2–Al2O3 catalyst at 473 K for 6 h was 31.7%, which was higher than
O3 catalyst at 473 Kfrom
yield obtained for
6the
h was
first31.7%, which
reaction. Onewas higher reason
possible than the is yield obtained
that the from the
calcination first10%ZrO
of the reaction. One possible reason is
2–Al2O3 catalyst after the
that
reaction would eliminate contamination on the surface, resulting in the higher lactic contamination
the calcination of the 10%ZrO 2 –Al 2 O 3 catalyst after the reaction would eliminate acid yield. This
on the surface, resulting in the higher lactic acid yield.
result suggests that the 10%ZrO2–Al2O3 catalyst is recyclable. This result suggests that the 10%ZrO 2 –Al2 O3
catalyst is recyclable.
2.4. Characterization
2.4. Characterization
Next, we characterized the ZrO2–Al2O3 catalysts to further understand their catalytic activity.
Next, we characterized the ZrO2 –Al2 O3 catalysts to further understand their catalytic activity. The
The X‐ray diffraction (XRD) patterns of the catalysts are shown in Figure 5. In the XRD patterns of
X-ray diffraction (XRD) patterns of the catalysts are shown in Figure 5. In the XRD patterns of the three
the three ZrO2–Al2O3 catalysts, peaks at 46.0° and 66.5° were attributed to γ‐Al2O3 [22]. In the XRD
ZrO2 –Al2 O3 catalysts, peaks at 46.0◦ and 66.5◦ were attributed to γ-Al2 O3 [22]. In the XRD patterns of
patterns of 10%ZrO2–Al2O3 and 20%ZrO2–Al2O◦3, peaks ◦at 30.3° and 50.7° were attributed to
10%ZrO2 –Al2 O3 and 20%ZrO2 –Al2 O3 , peaks at 30.3 and 50.7 were attributed to tetragonal ZrO2 [23];
tetragonal ZrO2 [23]; however, these peaks were not seen in the XRD pattern of 5%ZrO2–Al2O3,
however, these peaks were not seen in the XRD pattern of 5%ZrO2 –Al2 O3 , indicating that ZrO2 was
indicating that ZrO2 was either highly dispersed or in the amorphous phase in this catalyst. The
either highly dispersed or in the amorphous phase in this catalyst. The XRD pattern for pure ZrO2
XRD pattern for pure ZrO2 revealed that it was in the monoclinic phase (Figure 5d).
revealed that it was in the monoclinic phase (Figure 5d).
In our previous paper, we hypothesized that the properties of the acid and base sites on ZrO2
In our previous paper, we hypothesized that the properties of the acid and base sites on ZrO2
play an important role in the catalytic activity of this oxide for the conversion of cellulose to lactic
play an important role in the catalytic activity of this oxide for the conversion of cellulose to lactic
acid [17]. Therefore, we examined the acid and base properties of the ZrO2–Al2O3 catalysts by means
acid [17]. Therefore, we examined the acid and base properties of the ZrO2 –Al2 O3 catalysts by
of temperature‐programmed desorption of ammonia (NH3‐TPD) and temperature‐programmed
means of temperature-programmed desorption of ammonia (NH3 -TPD) and temperature-programmed
desorption of carbon dioxide (CO2‐TPD), respectively. Figure 6 shows NH3‐TPD profiles of the
desorption of carbon dioxide (CO2 -TPD), respectively. Figure 6 shows NH3 -TPD profiles of the
ZrO2 –Al2 O3 and ZrO2 catalysts. The NH3 desorption temperatures of the three ZrO2 –Al2 O3 catalysts
Catalysts 2017, 7, 221 5 of 10

were all 470 K, 2017,


Catalysts which was similar to that of ZrO2 , indicating that the acid sites on the ZrO2 –Al
7, 221 2 O3 and
5 of 10
ZrO2 catalysts were weakly acidic Lewis acid sites [24] (Figure 6). The shoulder peaks at 550 K were
ZrO2–Al
Catalysts 2O
2017, and ZrO2 catalysts. The NH3 desorption temperatures of the three ZrO2–Al2O3 catalysts
7,3 221 5 of 10
observed in the case of ZrO –Al O3 catalysts, which were ascribed to slightly stronger acid sites than
were all 470 K, which2was 2similar to that of ZrO2, indicating that the acid sites on the ZrO2–Al2O3
weaklyZrOacid
and2–Al
sites
ZrO (peak
2O23 catalysts
and ZrO2were at 470 K).
catalysts.
weakly The slightly
Theacidic
NH3 Lewis stronger
desorption acid
temperatures
acid sites sites
[24] (Figure might
of the be located
threeshoulder
6). The ZrO2–Alpeaks on the
2O3 catalysts
boundary
at 550 K
between ZrO
were
were and
all2observed
470 K, Alwhich
O3the
2in . Although
was
casesimilar
of ZrO the
to2–AlXRD
that 3 analysis
2Oof revealed
ZrO2, indicating
catalysts, which were thatthat
the the
acidcrystal
ascribed sites onphase
to slightly of2–Al
the stronger
ZrO ZrO 2O 23 differed
acid
depending
and onthan
ZrO
sites whether
weaklyitwere
2 catalysts wassites
acid in pure
weakly (peak orat mixed-oxide
acidic Lewis
470 K). acid
The sitesform, the
[24]stronger
slightly NHacid
(Figure 3 desorption
6). Thesites mighttemperatures
shoulder peaks
be locatedat 550
on K revealed
the
that thewere observed
boundary
acid strength inwas
between theZrO case2 and
almost of ZrO Al22O
the –Al 2O3irrespective
catalysts,
3. Although
same the which
XRD were ascribed
ofanalysis
form. revealedtothat slightly stronger
the crystal acidof
phase
sites
ZrOthan weaklydepending
acid sites (peak
The 2 differed
desorption temperatures ofatCO
on whether 470 K). The slightly
it was in pure stronger acid sitesform,
or mixed‐oxide mightthe
2 from the ZrO2 –Al2 O3 and ZrO2 catalysts were in the range
be NH
located on the
3 desorption

boundary
temperatures between ZrO2 that
revealed andthe Al2acid
O3. Although
strength was the almost
XRD analysis
the same revealed
irrespective that the crystal phase of
of form.
340–450ZrO K2 (Figure
differed
7), indicating
depending
that theit base
on whether
sites on the ZrO2 catalysts were 3 weakly basic [25].
The desorption temperatures of CO2was from inthe
pureZrOor mixed‐oxide
2–Al 2O3 and ZrO form, the NH
2 catalysts were desorption
in the range
The amount of
temperatures CO
340–450 K (Figure desorbed
2revealed 7), that from
the acid
indicating the
thatZrO –Al
the2base
strength was O
2sites catalysts
3 on the
almost theZrO was
same smaller
irrespective
2 catalysts
than
wereofweakly that desorbed
form. basic [25]. The from the
ZrO2 catalyst.The desorption
amount of CO2 desorbed temperatures
from theofZrO CO22–Alfrom2O3 the ZrO2–Al
catalysts was 2Osmaller
3 and ZrO than 2 catalysts
that desorbed were in fromthethe
range
ZrO2
340–450
Tablecatalyst. K (Figure
1 shows the7),amounts
indicating that the base
of acid andsites on the
base sitesZrOon 2 catalysts
the ZrO were 2 weakly
–Al O
2 3 basic
and [25].
ZrO The
2 catalysts.
The ZrO amount
–Al of
Table
O CO desorbed
12 shows
catalysts the
hadfrom
amounts
morethe ZrOacid –Al
of 2acid 2O
and
sites 3 catalysts
basefar
but was
sites
fewer smaller
on the
baseZrOthan2–Al
sites that2O3desorbed
than and theZrO from
ZrO the
2 catalysts.ZrO
catalyst.The
2
In our
2 2 3 2
catalyst.
ZrO2–Al2O3 catalysts had more acid sites but far fewer base sites than the ZrO2 catalyst. In our
previous paper, we suggested that the retro-aldol reaction (the conversion of fructose to glyceraldehyde
Table paper,
previous 1 shows wethe amounts
suggested of the
that acid and basereaction
retro‐aldol sites on(the theconversion
ZrO2–Al2Oof 3 and ZrOto
fructose 2 catalysts. The
glyceraldehyde
and dihydroxyacetone)—the
ZrO
and2–Al 2O3 catalysts had morekey
dihydroxyacetone)—the
key step insites
acidstep the conversion
in but
the far fewer of
conversion
cellulose
base sites than
of cellulose
to lactic
to the ZrO
lactic
acid—involved
2 catalyst. In our
acid—involved both
both the
acid and base
previous
the acid sites
andon
paper, wethe
base ZrO
suggested
sites catalyst
on2 thethatZrO [17]. However,
the2 retro‐aldol
catalyst reaction
[17]. However,the ZrO 2 –Al
(the conversion
the ZrO O23fructose
2of
2–Al Ocatalysts had moreacid
to glyceraldehyde
3 catalysts had more acid sites
and farand
fewer
sites base
and farsites
fewerthan
dihydroxyacetone)—the didkey
base sites ZrO
thanstep. We
2did in
ZrOdiscuss
the the reaction
conversion
2. We discuss pathway
of cellulose
the reaction to lactic
pathway inacid—involved
inthethenext
next section. both
section.
the acid and base sites on the ZrO2 catalyst [17]. However, the ZrO2–Al2O3 catalysts had more acid
sites and far fewer base sites than did ZrO2. We discuss the reaction pathway in the next section.

Figure 5. X‐ray diffraction patterns of (a) 5%ZrO2–Al2O3, (b) 10%ZrO2–Al2O3, (c) 20%ZrO2–Al2O3,
Figure 5. X-ray diffraction patterns of (a) 5%ZrO2 –Al2 O3 , (b) 10%ZrO2 –Al2 O3 , (c) 20%ZrO2 –Al2 O3 ,
and (d) ZrO2.
and (d)Figure
ZrO2 .5. X‐ray diffraction patterns of (a) 5%ZrO2–Al2O3, (b) 10%ZrO2–Al2O3, (c) 20%ZrO2–Al2O3,
and (d) ZrO2.

Figure 6. Temperature-programmed desorption of ammonia profiles of (a) 5%ZrO2 –Al2 O3 ,


(b) 10%ZrO2 –Al2 O3 , (c) 20%ZrO2 –Al2 O3 , and (d) ZrO2 .
Catalysts 2017, 7, 221 6 of 10

Figure 6. Temperature‐programmed desorption of ammonia profiles of (a) 5%ZrO2–Al2O3,


Catalysts 2017,(b)
7, 10%ZrO
221 2–Al2O3, (c) 20%ZrO2–Al2O3, and (d) ZrO2.
6 of 10

FigureFigure 7. Temperature‐programmed desorption


7. Temperature-programmed desorption ofofcarbon dioxide
carbon profiles
dioxide of (a)of5%ZrO
profiles 2–Al2O3, –Al O ,
(a) 5%ZrO 2 2 3
(b) 10%ZrO 2–Al2O3, (c) 20%ZrO2–Al2O3, and (d) ZrO2.
(b) 10%ZrO2 –Al2 O3 , (c) 20%ZrO2 –Al2 O3 , and (d) ZrO2 .

Table 1. Properties of the ZrO2–Al2O3 and ZrO2 catalysts.


Table 1. Properties of the ZrO2 –Al2 O3 and ZrO2 catalysts.
Surface Crystal Phase Amount of Acid Amount of Base Yield of Lactic
Catalyst
Area 1/m2 g−1 of ZrO2 2 Sites 3/mmol g−1 Sites 4/mmol g−1 Acid 5/%
5%ZrO2–Al2O3 Surface 135 Crystal
‐ Phase Amount
0.109 of Acid Amount of Base 25.0 Yield of Lactic
0.093
Catalyst 1 /m2 g−1 3 /mmol g−1 −1
10%ZrO2–Al2OArea
3 126 of ZrO2 2
Tetragonal Sites0.131 Sites 4
0.057 /mmol g 25.3 Acid /%
5

20%ZrO2–Al2O3 130 Tetragonal 0.180 0.098 21.9


5%ZrO2 –Al2 O3 6 135 - 0.109 0.093 25.0
ZrO2 101 Monoclinic 0.118 0.485 21.2
10%ZrO2 –Al2 O3 126 Tetragonal 0.131 0.057 25.3
20%ZrOReported by manufacturer;
1 2 Determined by XRD; 3 Determined by temperature‐programmed
2 –Al2 O3 130 Tetragonal 0.180 0.098 21.9
ZrO 6
desorption of ammonia
101 (NH3‐TPD); 4 Determined by temperature‐programmed desorption of
Monoclinic 0.118 0.485 21.2
2
carbon dioxide (CO2‐TPD);
1 Reported
5 Reaction conditions: ball‐milled cellulose, 0.5 g; catalyst, 1.0 g; water,
by manufacturer; 2 Determined by XRD; 3 Determined by temperature-programmed desorption
50 g; reaction
of ammonia temperature,
(NH3 -TPD); 473 K; reaction
4 Determined time, 6 h; 6 Presented in desorption
by temperature-programmed reference [17].
of carbon dioxide (CO2 -TPD);
5 Reaction conditions: ball-milled cellulose, 0.5 g; catalyst, 1.0 g; water, 50 g; reaction temperature, 473 K; reaction
time, h; 6 Presented
2.5.6Reaction Pathway
in reference [17].

The conversion of cellulose to lactic acid is a multi‐step reaction (Scheme 1). First, cellulose is
2.5. Reaction Pathway
converted to glucose by hydrolysis, which can be catalyzed by acid catalysts [26,27]. Then, glucose
is isomerized to fructose by a Lewis acid or basic catalysts [28]. The key step in the conversion
The conversion of cellulose to lactic acid is a multi-step reaction (Scheme 1). First, cellulose is
pathway from cellulose to lactic acid is the conversion of fructose to glyceraldehyde and
converted to glucose byvia
dihydroxyacetone hydrolysis,
a retro‐aldolwhich can
reaction be catalyzed
involving C–C bondby acid catalysts
cleavage, which[26,27]. Then, glucose is
can be enhanced
isomerized
by Lewis acids [29,30]. In a previous paper, we hypothesized that the combination of acid and basepathway
to fructose by a Lewis acid or basic catalysts [28]. The key step in the conversion
sites on ZrO
from cellulose 2 enhanced
to lactic acid isthethe
conversion
conversion of fructose to glyceraldehyde
of fructose to glyceraldehydeand dihydroxyacetone [17].
and dihydroxyacetone via
However,
a retro-aldol in the present
reaction involvingstudy,
C–Cthe bond
yield of lactic acidwhich
cleavage, obtainedcanfrom
be cellulose
enhanced by by
using the ZrO
Lewis 2–
acids [29,30].
Al2O3 catalysts did not appear to depend on the relative amount of acid and base sites. That is, the
In a previous paper, we hypothesized that the combination of acid and base sites on ZrO2 enhanced
ZrO2–Al2O3 catalysts provided a higher yield of lactic acid than did the ZrO2 catalyst, even though
the conversion of fructose to glyceraldehyde and dihydroxyacetone [17]. However, in the present
they had more acid sites and far fewer base sites than did ZrO2 (Table 1). This indicates that the
study, number
the yield of lactic
of base acid
sites is not obtained
important from
for thecellulose by using of
catalytic conversion thecellulose
ZrO2 –Al 2 O3 catalysts
to lactic acid with did not
appearZrOto 2depend on the relative
–Al2O3 catalysts. amount
Furthermore, of acid
although the and
Lewisbase
acidsites.
sites That
on the is,ZrO
the2–Al
ZrO 2O23–Al 2 O3 catalysts
catalysts
played
provided an important
a higher yield role in the acid
of lactic conversion
than of did cellulose
the ZrOto lactic
2 acid,
catalyst, the yield
even of lactic
though acid
they obtained
had more acid
wasfar
sites and notfewer
proportional to thethan
base sites number
didofZrOacid sites.
(TableThis
1).suggests
This that dispersion
indicates that of ZrO
the 2 on a Al2O3
number of base sites
2
support and accessibility to the acid sites on the catalyst are the most important factors for the
is not important for the catalytic conversion of cellulose to lactic acid with ZrO2 –Al2 O3 catalysts.
conversion of cellulose to lactic acid when using ZrO2–Al2O3 catalysts. Further studies are required
Furthermore, although the Lewis acid sites on the ZrO2 –Al2 O3 catalysts played an important role in
the conversion of cellulose to lactic acid, the yield of lactic acid obtained was not proportional to the
number of acid sites. This suggests that dispersion of ZrO2 on a Al2 O3 support and accessibility to
the acid sites on the catalyst are the most important factors for the conversion of cellulose to lactic
acid when using ZrO2 –Al2 O3 catalysts. Further studies are required to elucidate the details of the
relationship between the active sites and the product yield and also the reaction mechanism.
Catalysts 2017, 7, 221 7 of 10

to elucidate
Catalysts 2017, 7,the
details of the relationship between the active sites and the product yield and7 of
221 also
10
the reaction mechanism.

Scheme 1.
Scheme 1. Reaction
Reaction pathway
pathway for
for the
the conversion
conversion of
of cellulose
cellulose to
to lactic
lactic acid.
acid.

In the final step of the reaction pathway, glyceraldehyde and dihydroxyacetone can be
In the final step of the reaction pathway, glyceraldehyde and dihydroxyacetone can be converted
converted to lactic acid at 463 K without a catalyst [17]. However, the conversion of pyruvaldehyde
to lactic acid at 463 K without a catalyst [17]. However, the conversion of pyruvaldehyde to lactic acid is
to lactic acid is reported to require a catalyst with Lewis acid sites [31,32], which a ZrO2‐based
reported to require a catalyst with Lewis acid sites [31,32], which a ZrO2 -based catalyst could provide.
catalyst could provide.
Previously, we reported that we used a ZrO2 catalyst to convert cellulose to lactic acid with
Previously, we reported that we used a ZrO2 catalyst to convert cellulose to lactic acid with a
a yield of 21.2% [17]. However, this yield of lactic acid was less than that obtained with other reported
yield of 21.2% [17]. However, this yield of lactic acid was less than that obtained with other
heterogeneous catalysts (24% yield of lactic acid using LaCoO3 [14], 28% using AlW [16], and 27%
reported heterogeneous catalysts (24% yield of lactic acid using LaCoO3 [14], 28% using AlW [16],
using NbF5 –AlF3 [15]). In the present study, we found that by using the 10%ZrO2 –Al2 O3 catalyst we
and 27% using NbF5–AlF3 [15]). In the present study, we found that by using the 10%ZrO2–Al2O3
could obtain a yield of lactic acid of 25.3%, which was comparable with the yields using the reported
catalyst we could obtain a yield of lactic acid of 25.3%, which was comparable with the yields using
catalysts. One advantage of the 10%ZrO2 –Al2 O3 catalyst is that it can be obtained in lower cost than
the reported catalysts. One advantage of the 10%ZrO2–Al2O3 catalyst is that it can be obtained in
the reported heterogeneous catalysts. Additionally, a few percentages of the metal species in the
lower cost than the reported heterogeneous catalysts. Additionally, a few percentages of the metal
reported catalysts leached into solution during the reaction. Conversely, the Zr species was not leached
species in the reported catalysts leached into solution during the reaction. Conversely, the Zr
out from ZrO2 into water during the reaction [17].
species was not leached out from ZrO2 into water during the reaction [17].
3. Materials and Methods
3. Materials and Methods
3.1. Materials
3.1. Materials
ZrO2 (ZRO-7; reference catalyst from Catalysis Society of Japan) and mixed oxides containing
ZrO2 ZrO
and2Al(ZRO‐7; reference
2 O3 (5%ZrO 2 –Al2catalyst from2 –Al
O3 , 10%ZrO Catalysis Society
2 O3 , and 20%ZrOof 2Japan)
–Al2 O3and mixed
) were oxides
obtained containing
from Daiichi
ZrO2 andKagaku
Kigenso Al2O3 (5%ZrO
Kogyo 2–Al
Co.,2OLtd.
3, 10%ZrO
(Osaka,2–Al 2O3, and
Japan). 20%ZrO2–Al
Aluminum 2O3) (Al
oxide were O
2 3 obtained
) was from
obtainedDaiichi
from
Kigenso Kagaku
Sigma–Aldrich Co.,Kogyo Co.,
LLC (St. Ltd. MO,
Louis, (Osaka,
USA).Japan).
These Aluminum
materials were oxide
used (Al 2O3) was
without obtained from
pretreatment.
Sigma–Aldrich Co., LLC (St. Louis, MO, USA). These materials were used without pretreatment.
3.2. Catalytic Reaction
3.2. Catalytic Reaction
Cellulose (microcrystalline cellulose; Merck KGaA, Darmstadt, Germany) was pulverized
with Cellulose
a ball mill at 60 rpm. cellulose;
(microcrystalline The conversion of cellulose
Merck KGaA, to lactic
Darmstadt, acid was pulverized
Germany) carried outwithin
aa stainless 3
ball mill steel
at 60 batch
rpm. reactor with an of
The conversion inner volume
cellulose to of 100 acid
lactic cm was (MMJ-100;
carriedOMout Lab-Tech, Tochigi,
in a stainless steel
Japan), as described in a previous paper [17]. Briefly, the reactor was loaded with ball-milled
Catalysts 2017, 7, 221 8 of 10

cellulose (0.5 g), catalyst (1.0 g), and water (50 g); purged with nitrogen gas (0.1 MPa); and
then heated to 453–493 K for 3–9 h with screw stirring. After the reaction was allowed to
run, the resulting mixture was filtered to separate the liquid from the solid. The quantitative
analyses of lactic acid, levulinic acid, 5-hydroxymethylfurfural, and furfural in the liquid fractions
were carried out with a gas chromatograph (GC-2014; Shimadzu, Kyoto, Japan) equipped with
a flame ionization detector and an InertCap capillary column (GL Sciences Inc., Tokyo, Japan).
The chemicals in the liquid fraction (e.g., glucose) were analyzed by using a high-performance
liquid chromatograph (Shimadzu, Kyoto, Japan) equipped with a refractive index detector (RID-10A;
Shimadzu, Kyoto, Japan), an ultraviolet/ visible detector (SPD-20AV; Shimadzu, Kyoto, Japan),
and a Rezex RPM-Monosaccharide Pb+2 column (Phenomenex Inc., Torrance, CA, USA). The product
yields were calculated based on moles of carbon as follows:

(Moles of carbon atoms in each product)


Product yield (%) = × 100 (1)
(Moles of carbon atoms in cellulose used)

(Weight of solid residue) − (Weight of solid catalyst)


 
Conversion (%) = 1− × 100 (2)
(weight of cellulose used)

3.3. Catalyst Characterization


XRD patterns of the catalysts were determined by using a Rigaku SmartLab XRD system
(Rigaku, Tokyo, Japan) with Cu Kα radiation in the 2θ range of 5–90◦ .
NH3 -TPD profiles were determined with a TPD-1-AT instrument (Bel Japan, Inc., Osaka, Japan)
with an online quadrupole mass spectrometer. Samples (ca. 0.05 g) were pretreated at 773 K in flowing
helium for 1 h, saturated in flowing 5% ammonia diluted with helium (0.5 cm3 s−1 ) at 373 K for 30 min,
and then treated at 373 K in flowing helium (0.83 cm3 s−1 ) for 1 h. The samples were then heated at
a constant rate of 10 K min−1 from 373 to 953 K while the amount of NH3 desorbed was detected.
CO2 -TPD profiles were determined with a Micromeritics 3FLEX 3500 chemisorption analyzer
(Micromeritics, Norcross, GA, USA) with an online quadrupole mass spectrometer. Samples (ca. 0.2 g)
were pretreated at 773 K in flowing helium for 1 h, saturated in flowing CO2 (0.5 cm3 s−1 ) at 323 K
for 30 min, and then treated at 323 K in flowing helium (0.83 cm3 s−1 ) for 1 h. The samples were
then heated at a constant rate of 10 K min−1 from 323 to 953 K while the amount of CO2 desorbed
was detected.

4. Conclusions
Compared with using pure ZrO2 as the catalyst (yield, 21.2%), a greater yield of lactic acid
was obtained from cellulose by using a 10%ZrO2 –Al2 O3 catalyst (yield, 25.3%) and optimized
reaction conditions (reaction temperature, 473 K; reaction time, 6 h). XRD analysis revealed that the
10%ZrO2 –Al2 O3 catalyst contained tetragonal ZrO2 and γ-Al2 O3 . NH3 -TPD and CO2 -TPD analyses
revealed that all three ZrO2 –Al2 O3 catalysts had more Lewis acid sites and far fewer base sites than
did ZrO2 . This suggests that the number of Lewis acid sites on the ZrO2 –Al2 O3 catalysts was more
important than the number of base sites for the conversion of cellulose to lactic acid.

Acknowledgments: We acknowledge Daiichi Kigenso Kagaku Kogyo Co., Ltd. (Osaka, Japan) for providing the
ZrO2 –Al2 O3 samples. This study was partially supported by a JSPS KAKENHI grant (JP17H00803), the Thailand
Research Fund (IRG5780001), and an NRCT-NSFC joint funding project (NRCT/2558-104).
Author Contributions: P.W. and A.Y. conceived and designed the experiments; P.W. performed the catalytic
reaction experiments; O.S. and N.M. analyzed the data; K.S. performed the NH3 -TPD study; P.W., P.R., and A.Y.
wrote the paper; all the authors discussed the results and commented on the manuscript.
Conflicts of Interest: The authors declare no conflict of interest.
Catalysts 2017, 7, 221 9 of 10

References
1. Huber, G.W.; Iborra, S.; Corma, A. Synthesis of Transportation Fuels from Biomass: Chemistry, Catalysts,
and Engineering. Chem. Rev. 2006, 106, 4044–4098. [CrossRef] [PubMed]
2. Petrus, L.; Noordermeer, M.A. Biomass to biofuels, a chemical perspective. Green Chem. 2006, 8, 861–867.
[CrossRef]
3. Corma, A.; Iborra, S.; Velty, A. Chemical Routes for the Transformation of Biomass into Chemicals. Chem. Rev.
2007, 107, 2411–2502. [CrossRef] [PubMed]
4. Kobayashi, H.; Komanoya, T.; Guha, S.K.; Hara, K.; Fukuoka, A. Conversion of cellulose into renewable
chemicals by supported metal catalysis. Appl. Catal. A 2011, 409–410, 13–20. [CrossRef]
5. Gallezot, P. Conversion of biomass to selected chemical products. Chem. Soc. Rev. 2012, 41, 1538–1558.
[CrossRef] [PubMed]
6. Yamaguchi, A. Biomass Valorization in High-temperature Liquid Water. J. Jpn. Petrol. Inst. 2014, 57, 155–163.
[CrossRef]
7. Yamaguchi, A.; Sato, O.; Mimura, N.; Shirai, M. One-pot conversion of cellulose to isosorbide using supported
metal catalysts and ion-exchange resin. Catal. Commun. 2015, 67, 59–63. [CrossRef]
8. Datta, R.; Henry, M. Lactic acid: recent advances in products, processes and technologies—A review. J. Chem.
Technol. Biotechnol. 2006, 81, 1119–1129. [CrossRef]
9. Mäki-Arvela, P.; Simakova, I.L.; Salmi, T.; Murzin, D.Y. Production of Lactic Acid/Lactates from Biomass
and Their Catalytic Transformations to Commodities. Chem. Rev. 2014, 114, 1909–1971. [CrossRef] [PubMed]
10. Garlotta, D. A Literature Review of Poly(Lactic Acid). J. Polym. Environ. 2001, 9, 63–84. [CrossRef]
11. Martinez, F.A.C.; Balciunas, E.M.; Salgado, J.M.; González, J.M.D.; Converti, A.; Oliveira, R.P.D.S. Lactic acid
properties, applications and production: A review. Trends Food Sci. Technol. 2013, 30, 70–83. [CrossRef]
12. Wang, Y.; Deng, W.; Wang, B.; Zhang, Q.; Wan, X.; Tang, Z.; Wang, Y.; Zhu, C.; Cao, Z.; Wang, G.; et al.
Chemical synthesis of lactic acid from cellulose catalysed by lead(II) ions in water. Nat. Commun. 2013, 4,
2141. [CrossRef] [PubMed]
13. Lei, X.; Wang, F.-F.; Liu, C.-L.; Yang, R.-Z.; Dong, W.-S. One-pot catalytic conversion of carbohydrate biomass
to lactic acid using an ErCl3 catalyst. Appl. Catal. A 2014, 482, 78–83. [CrossRef]
14. Yang, X.; Yang, L.; Fan, W.; Lin, H. Effect of redox properties of LaCoO3 perovskite catalyst on production of
lactic acid from cellulosic biomass. Catal. Today 2016, 269, 56–64. [CrossRef]
15. Coman, S.M.; Verziu, M.; Tirsoaga, A.; Jurca, B.; Teodorescu, C.; Kuncser, V.; Parvulescu, V.I.; Scholz, G.;
Kemnitz, E. NbF5 –AlF3 Catalysts: Design, Synthesis, and Application in Lactic Acid Synthesis from Cellulose.
ACS Catal. 2015, 5, 3013–3026. [CrossRef]
16. Chambon, F.; Rataboul, F.; Pinel, C.; Cabiac, A.; Guillon, E.; Essayem, N. Cellulose hydrothermal conversion
promoted by heterogeneous Brønsted and Lewis acids: Remarkable efficiency of solid Lewis acids to produce
lactic acid. Appl. Catal. B 2011, 105, 171–181. [CrossRef]
17. Wattanapaphawong, P.; Reubroycharoen, P.; Yamaguchi, A. Conversion of cellulose into lactic acid using
zirconium oxide catalysts. RSC Adv. 2017, 7, 18561–18568. [CrossRef]
18. Teramoto, Y.; Tanaka, N.; Lee, S.-H.; Endo, T. Pretreatment of eucalyptus wood chips for enzymatic
saccharification using combined sulfuric acid-free ethanol cooking and ball milling. Biotechnol. Bioeng.
2008, 99, 75–85. [CrossRef] [PubMed]
19. Yamaguchi, A.; Hiyoshi, N.; Sato, O.; Bando, K.K.; Shirai, M. Gaseous Fuel Production from Nonrecyclable
Paper Wastes Using Supported Metal Catalysts in High-Temperature Liquid Water. ChemSusChem 2010, 3,
737–741. [CrossRef] [PubMed]
20. Yamaguchi, A.; Sato, O.; Mimura, N.; Hirosaki, Y.; Kobayashi, H.; Fukuoka, A.; Shirai, M. Direct Production
of Sugar Alcohols from Wood Chips using Supported Platinum Catalysts in Water. Catal. Commun. 2014, 54,
22–26. [CrossRef]
21. Zhao, H.; Kwak, J.H.; Wang, Y.; Franz, J.A.; White, J.M.; Holladay, J.E. Effects of Crystallinity on Dilute Acid
Hydrolysis of Cellulose by Cellulose Ball-Milling Study. Energy Fuels 2006, 20, 807–811. [CrossRef]
22. Said, A.E.-A.A.; El-Wahab, M.M.M.A.; El-Aal, M.A. Effect of ZrO2 on the catalytic performance of nano
γ-Al2 O3 in dehydration of methanol to dimethyl ether at relatively low temperature. Res. Chem. Intermed.
2016, 42, 1537–1556. [CrossRef]
Catalysts 2017, 7, 221 10 of 10

23. Zhang, D.; Duan, A.; Zhao, Z.; Wan, G.; Gao, Z.; Jiang, G.; Chi, K.; Chuang, K.H. Preparation, characterization
and hydrotreating performances of ZrO2 –Al2 O3 -supported NiMo catalysts. Catal. Today 2010, 149, 62–68.
[CrossRef]
24. Manríquez, M.E.; López, T.; Gómez, R.; Navarrete, J. Preparation of TiO2 –ZrO2 mixed oxides with controlled
acid–basic properties. J. Mol. Catal. A 2004, 220, 229–237. [CrossRef]
25. Ma, Z.-Y.; Yang, C.; Wei, W.; Li, W.-H.; Sun, Y.-H. Surface properties and CO adsorption on zirconia
polymorphs. J. Mol. Catal. A 2005, 227, 119–124. [CrossRef]
26. Rinaldi, R.; Schuth, F. Acid Hydrolysis of Cellulose as the Entry Point into Biorefinery Schemes. ChemSusChem
2009, 2, 1096–1107. [CrossRef] [PubMed]
27. Wang, J.; Xi, J.; Wang, Y. Recent advances in the catalytic production of glucose from lignocellulosic biomass.
Green Chem. 2015, 17, 737–751. [CrossRef]
28. Delidovich, I.; Palkovits, R. Catalytic Isomerization of Biomass-Derived Aldoses: A Review. ChemSusChem
2016, 9, 547–561. [CrossRef] [PubMed]
29. Yang, L.; Yang, X.; Tian, E.; Lin, H. Direct Conversion of Cellulose into Ethyl Lactate in Supercritical
Ethanol–Water Solutions. ChemSusChem 2016, 9, 36–41. [CrossRef] [PubMed]
30. Yang, L.; Yang, X.; Tian, E.; Vattipalli, V.; Fan, W.; Lin, H. Mechanistic insights into the production of methyl
lactate by catalytic conversion of carbohydrates on mesoporous Zr-SBA-15. J. Catal. 2016, 333, 207–216.
[CrossRef]
31. Pescarmona, P.P.; Janssen, K.P.F.; Delaet, C.; Stroobants, C.; Houthoofd, K.; Philippaerts, A.; De Jonghe, C.;
Paul, J.S.; Jacobs, P.A.; Sels, B.F. Zeolite-catalysed conversion of C3 sugars to alkyl lactates. Green Chem. 2010,
12, 1083–1089. [CrossRef]
32. Nakajima, K.; Noma, R.; Kitano, M.; Hara, M. Titania as an Early Transition Metal Oxide with a High Density
of Lewis Acid Sites Workable in Water. J. Phys. Chem. C 2013, 117, 16028–16033. [CrossRef]

© 2017 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).

You might also like