You are on page 1of 9

Molecular Catalysis 525 (2022) 112353

Contents lists available at ScienceDirect

Molecular Catalysis
journal homepage: www.journals.elsevier.com/molecular-catalysis

High selective oxidation of 5-hydroxymethyl furfural to 5-hydroxyme­


thyl-2-furan carboxylic acid using Ag-TiO2
Tao Xie a, 1, Shitong Yue a, 1, Ting Su b, *, Mingqi Song a, Wenjie Xu a, Yaxi Xiao a,
Zhenglong Yang a, Christophe Len c, d, **, Deyang Zhao a, *
a
School of Chemistry and Materials Science, Ludong University, Yantai 264025, China
b
Green Chemistry Centre, College of Chemistry and Chemical Engineering, Yantai University, Yantai 264005, China
c
Chimie ParisTech, CNRS, Institute of Chemistry for Life and Health Sciences, PSL Research University, 11 rue Pierre et Marie Curie, Paris F-75005, France
d
Université de Technologie de Compiègne, Compiegne F-60200, France

A R T I C L E I N F O A B S T R A C T

Keywords: Production of 5-hydroxymethyl-2-furan carboxylic acid (HMFCA) from 5-hydroxymethylfurfural (HMF) selective
HMF oxidation using Ag-TiO2 were investigated under oxygen atmosphere. The novel catalyst was prepared by a
Selective oxidation simple sol-gel method, different characterizations were performed, which included N2 adsorption and desorption
HMFCA
measurements, X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), NH3/CO2 temperature-
Ag-TiO2
programmed desorption (NH3/CO2-TPD), scanning electron microscope (SEM), and high revolution trans­
mission electron microscope (HR-TEM). Reaction parameters, such as temperature, base content, oxygen pres­
sure, catalyst loading and reaction time were investigated. 99% HMF conversion, 96% HMFCA yield, and 97%
HMFCA selectivity obtained under optimum condition (25 mg Ag-TiO2, 4 eq. Na2CO3, 1 MPa O2, 80 ◦ C, 32 h).
Catalyst stability experiment revealed negligible deactivation (90% HMF conversion and 85% HMFCA yield in
fourth time). The appropriate proportion of acidic/basic sites, large surface areas, and strong synergistic inter­
action between Ag+, metallic Ag and TiO2 for oxygen activation attributed to high catalytic activation.

1. Introduction and as the pharmacy sector it showed antitumor activity [8], as well as
excellent interleukin inhibition [9]. However, only a limited number of
Both drastic depletion of fossil resources and severe carbon emission studies focus on the high selectivity synthesis of HMFCA by applying
situation have stimulated considerable attention to exploring renewable various catalytic processes [10,11], which may be due to easy over
resources [1]. Biomass resources as the natural features of sustainability oxidation of HMF to FFCA or FDCA. Therefore, it is necessary to design
and environmental friendliness resources, which majority derived from an efficient catalyst for the high selectivity of HMF oxidation to HMFCA.
carbon dioxide fixed by plant photosynthesis [2]. Furfural (FF), TiO2 as a common material have been wildly investigated in biomass
5-hydroxymethylfurfural (HMF), which derived from the dehydration of valorization [12]. For instance, Riisager reported Au-TiO2 as catalyst for
xylose, glucose or fructose, are considered as the most versatile biomass HMFCA synthesis, 71% HMFCA yield achieved under optimum condi­
platform molecules [3]. Numerous attractive biomass based tion (Table 1, entry 1), the results depended on the amount of base and
value-added chemicals can obtain from FF [4,5] and HMF [6]. HMF the oxygen pressure. Further, Davis evaluated Au/TiO2+Au/C catalytic
oxidation intermediates including 2, 5-furanediformic acid (FDCA), 2, system under milder reaction condition with higher HMFCA produc­
5-diformylfuran (DFF), 5-hydroxymethyl-2-furanediformic acid tivity (24.1 molHMFCA/molmetal/h) (Table 1, entry 2). It proved HMFCA
(HMFCA), 5-formyl-2-furancarboxylic acid (FFCA) and so on (Scheme could not convert to FDCA at specified O2 pressure and base concen­
1), all of them have important applications. In particular, HMFCA rep­ tration. In contrast, FDCA was easily formed over Pt/C and Pd/C under
resents a novel monomer for the synthesis of furan-based polyesters [7], identical conditions. Ebitani disclosed another impressive result using

* Corresponding authors.
** Corresponding author at: Chimie ParisTech, CNRS, Institute of Chemistry for Life and Health Sciences, PSL Research University, 11 rue Pierre et Marie Curie,
Paris F-75005, France.
E-mail addresses: tingsu@ytu.edu.cn (T. Su), christophe.len@chimieparistech.psl.eu (C. Len), deyang.zhao@ldu.edu.cn (D. Zhao).
1
These authors contributed equally to this work.

https://doi.org/10.1016/j.mcat.2022.112353

Available online 11 May 2022


2468-8231/© 2022 Elsevier B.V. All rights reserved.
T. Xie et al. Molecular Catalysis 525 (2022) 112353

Table 1
Summary of HMFCA production under different catalytic system.
Entry Catalyst Operating conditions Catalytic properties Refs.
Temp.( ◦ C) Time NaOH Oxygen HMF Conv. Yield Productivity(molHMFCA/molmetal/
(h) (eq.) (bar) (M) (%) (%) h)

1 Au/TiO2 30 18 20 20 0.1 100 71 3.9 [13]


2 Au/TiO2 22 6 2 7 0.15 100 92 24.1 [14]
+Au/C
3 Au/HT 30 7 —- flow 0.16 100 87 4.9 [15]
50 mL/min
4 Au-Cu/TiO2 60 4 4 10 0.1 100 92 23.1 [16]
5 Au/ 65 1 4 10 0.15 100 64 95.5 [17]
Ce0.9Bi0.1Ox
6 Ru/CsPW 130 12 —- flow 0.06 97 73 1.8 [18]
20 mL/min
7 K-10 clay-Mo 110 3 —- flow 0.14 100 87 19.3 [19]
20 mL/min
8 Ag/ZrO2 50 5 25 10 (air) 0.1 100 98 400 [20]
9 Ag-TiO2 80 32 4 (Na2CO3) 10 0.01 99 96 42.1 This work

Au/HT as catalyst without the addition of any homogeneous base. 87% activation with good recyclability. However, the catalytic mechanism of
HMFCA achieved under O2 flow 50 mL/min (Table 1, entry 3). Bime­ Ag-TiO2 performing in HMF selective oxidation was rarely investigated.
tallic Au-Cu/TiO2 displayed superior activity (Table 1, entry 4) as Herein, a novel catalyst (Ag/TiO2) through a facile sol-gel hydrothermal
compared to monometallic Au/TiO2 and Cu/TiO2. Moreover, the method were explored and performed as catalyst under various reaction
bimetallic Au-Cu/TiO2 can be recovered by filtration and reused without conditions to obtain the higher catalytic activation (42.1 molHMFCA/­
significant loss of activity and selectivity than monometallic catalysts. molmetal/h HMFCA productivity in Table 1, entry 9).
Noticeably, Yang et al. investigated serious Au/Ce1− xBixO2− δ (0.08 < x
< 0.5), 95.5 HMFCA productivity achieved under optimum condition 2. Experimental
(Table 1, entry 5), which was attributed to oxygen activation and hy­
dride transfer enhanced by Bi doping and the large amount of oxygen 2.1. Materials
vacancies. Other noble metal, like Ru/CsPW were studied by Zhang
without the addition of base. 73% HMFCA yield obtained under opti­ HMF (≥ 99.5%), HMFCA (≥ 99.5%), titanium isopropoxide (≥ 95%),
mum condition (Table 1, entry 6). Unfortunately, a slight loss of cata­ n-decane (≥ 99%) and silver nitrate (AgNO3) were purchased from
lytic activity happened during the recycling experiments. The same Maclin company. Sodium carbonate (Na2CO3) purchased from Sigma-
research group performed a molybdenum (VI) complex immobilized on Aldrich. All the chemicals were used without purification.
K-10 clay as catalyst for HMFCA production, the non-noble Mo afforded
impressive 19.3 molHMFCA/molmetal/h productivity (Table 1, entry 7). 2.2. Catalyst preparation
More importantly, K-10 clay-Mo could be reused several times without a
significant loss of catalytic activation. The synthesis method of 5% Ag-TiO2 was identical with our previous
Ag-based catalysts have attracted particular attention due to their work [31,32]. Firstly, 10 mL titanium isopropoxide solution was added
high selectivity towards aldehydes to the corresponding carboxylic acids into a glass beaker with 60 mL deionized water. 0.3152 g AgNO3 was
[21]. For instance, Li reported homogeneous silver(I)-catalyzed aerobic added into the flask with strong stirring. The mixture was heated under
oxidation of more than 50 examples of different aliphatic and aromatic 90 ◦ C until gel left to 20 mL, then moved into a sealed Teflon-lined
aldehydes in water [22]. In our previous work, Ag2O, Ag2O–CaCO3 autoclave in the 200 ◦ C oven for 12 h. The sample was washed with
were investigated for HMF selective oxidation to HMFCA under micro­ deionized water and methanol for several times and dried under 80 ◦ C
wave heating method for homogeneous or heterogeneous base condi­ for 12 h. Finally, the sample was transferred to a muffle oven under
tions [23,24]. In addition, different mechanical pathways using DFT 500 ◦ C for 0.5 h to obtain the final catalyst, and labeled as 5% Ag-TiO2.
calculations to study the energy states of different intermediates for
HMFCA production through Ag2O have been investigated [25]. Subse­
2.3. Characterization methods
quently, Ag NPs supported on heterogeneous supports, such as
multi-wall CNTs [26] and Ag/ZrO2 [20] showed a remarkably high ac­
N2 adsorption and desorption isotherms measurement was con­
tivity in aldehyde oxidation producing HMFCA. The quantitative yields
ducted by a Micromeritics ASAP 2460 instrument. NH3/CO2
with a maximum productivity of 400 molHMFCAh− 1mol−Ag1 were obtained
temperature-programmed desorption (NH3/CO2-TPD) were performed
under Ag/ZrO2 (Table 1, entry 8). The oxygen activation mechanism
by AutoChem 2920 instrument (USA). X-ray diffraction (XRD Rigaku,
was investigated, it proved the oxygen molecules dissociated at the
Smart Lab3) was identified with a Cu Kα radiation (40 mA, 45 kV) in the
particle surfaces firstly and then the dissociated oxygen atoms diffused
range of 20–70º (2θ). X-ray photoelectron spectroscopy (XPS) data were
through Ag NPs to reach the Ag/support interfaces. However, Ag based
collected by an ESCALAB 250 electron spectrometer using 150 W Al-Kα
catalysts also have the following disadvantages: due to its inherent
radiations. Scanning electron microscopy (SEM) using a Rili SU
properties, aggregation can be generated by means of Ostwald ripening
8000HSD Series Hitachi. The high-resolution transmission electron mi­
especially at high temperatures. A stabilizer such as poly vinyl­
croscopy (HR-TEM) images performed by a JEOL 2100 microscope.
pyrrolidone (PVP) as a capping agent can effectively protect Ag from
aggregation [27]. Thus, catalytic process with advantages like easy re­
covery and recyclability are more desirable. 2.4. Experiments
Ag2O nanoparticles supported on TiO2 materials has been investi­
gated widely in photocatalyst field [28–30]. The synergistic effects and In a typical process, 37 mg HMF, 25 mg catalyst, 4 eq. Na2CO3, 30 mL
efficient charge separation efficiency increased the interaction between H2O in 50 mL micro reactor (Shanghai Yanzheng laboratory reactor),
Ag2O and TiO2, thus increase the dispersion and the photocatalytic the instrument was equipped with temperature control device, and set at
80 ◦ C. Oxygen was added in several times through the high-pressure

2
T. Xie et al. Molecular Catalysis 525 (2022) 112353

Fig. 1. N2 adsorption− desorption isotherms (a) and aperture distribution curve (b) of 5% Ag-TiO2 and TiO2.

3. Results and discussion


Table 2
Acid and base content, BET surface area, and pore size with different catalysts.
3.1. Catalyst characterization
Catalyst Acidity(cm3/ Basicity(cm3/ BET surface area Pore size
g)a g)a (m2/g)b (nm)b
Nitrogen physical adsorption-desorption of 5% Ag-TiO2 and pure
TiO2 6.81 9.42 36.94 20.07 TiO2 were recorded in Fig. 1. It can be seen that the isotherms of the two
5% Ag- 16.76 10.34 129.72 8.51
samples were type IV isotherms [33], indicating both TiO2 and 5%
TiO2
Ag-TiO2 prepared by the experiments were porous structural. Moreover,
a
Acid and base content were obtained from NH3 and CO2-TPD, respectively. compared to pure TiO2, the hysteresis ring of 5% Ag-TiO2 moved into
b
BET surface area and pore size were measured by N2 adsorption− desorption. relatively low pressure, indicating the aperture on TiO2 decreased after
Ag binding. The pore size distribution of isotherm adsorption branch of
oxygen cylinder to eliminate the air in the reactor and reach the pre­ each sample was calculated by BJH algorithm. The surface area of
determined pressure 1 MPa, reaction time 32 h under magnetic stirring Ag-TiO2 and pure TiO2 porous structure were studied using the
200 rpm. Brunauer-Emmett-Teller (BET) method, as shown in Table 2. It can be
easily found that the surface area of 5% Ag-TiO2 (129.72 m2/g) was
2.5. Products analysis larger than the surface area of TiO2 (36.94 m2/g), indicating TiO2
combined with Ag could adjust the structural network to achieve the
High performance liquid chromatography (HPLC), each sample was maximum porous surface area.
analyzed separately by photodiode array detectors with C-18 column Basicity and acidity property of 5% Ag-TiO2 were measured by CO2/
(Agilent Technologies 1260, USA). The mobile phase was acetonitrile NH3-TPD, the deconvolution peak fitting in the function of temperature
and water (1:9) solution flowing at a rate of 1 mL/min. The column oven exhibited in Fig. 2a and b. Five deconvolution peaks located around 258,
temperature and sample injection were set at 30 ◦ C, 2 μL. The retention 379, 472, 555 and 660 ◦ C were found in CO2-TPD (Fig. 2a), correspond
times for HMFCA, HMF were 7.6, 3.2 min, respectively. The precise to alkaline contents 2.78, 3.67, 2.39, 0.76 and 0.76 cm3/g, respectively.
element amount of Ag and Ti of the catalyst were analyzed by ICP-OES, According to the range of weak (25–200 ◦ C), medium (200–400 ◦ C) and
Ag element was 0.09% (wt%), Ti element was 28.4% (wt%). strong alkaline sites (> 400 ◦ C), the main alkaline source of 5% Ag-TiO2

Fig. 2. NH3/CO2-TPD spectrum of 5% Ag-TiO2.

3
T. Xie et al. Molecular Catalysis 525 (2022) 112353

located at 2θ = 25.41◦ , 30.88◦ , 37.88◦ , 48.05◦ , 54.87◦ , 62.71◦ , 69.45◦ ,


75.24◦ , which attributed to (110), (103), (200), (211), (204) and (116)
planes of anatase TiO2 (JCPDS NO.21–1272) [34]. The peak located at
2θ = 30.7◦ indicated a small fraction of brookite type in fresh TiO2
[JCPDS#29–1360]. For 5% Ag-TiO2 (black curve), due to the relatively
low content of Ag, no appearance of (111) and (200) facet of metallic
Ag0 at 38.0◦ and 44.1◦ were found [35]. Noticeably, Ag species doped
into TiO2 could be well distributed, without the formation of crystallo­
graphic planes of TiO2 anatase changed.
In order to confirm the conclusions from aforementioned charac­
terization, XPS analysis of C 1s, O 1s, Ti 2p, and Ag 3d were investigated
in Fig. 4. Meanwhile, the atomic concentration on the surface in XPS
section were recorded in Fig. S1. From the observation, O 1s, Ti 2p, C 1s
accounted for 50%, 23%, 21%, respectively. Ag 3d accounted for 2.3%,
as well as Si 2p. XPS region of C 1s exhibited three characteristic peaks
with contributions of 288.4 eV, 286.1 eV and 284.5 eV, corresponding to
COO− (ca. 288.0–288.5 eV), C–O (ca. 285.5 − 286.1 eV), and C–C/
C–– C (ca. 284.6 eV) [36], respectively. The presence of C– – O may be
Fig. 3. XRD pattern of TiO2 and 5% Ag-TiO2. caused by a small amount of CO2 adsorption in the air on catalyst sur­
face, and the appearance of C–O and C–C characteristic peaks related
was medium strength basicity (6.45 cm3/g) and strong strength basicity to a small amount of ethanol adsorbed on the surface of the catalyst
(3.91 cm3/g). Meanwhile, NH3-TPD of 5% Ag-TiO2 exhibited five during the preparation or contamination process.
deconvolution peaks, located at 110, 158, 239, 306 and 369 ◦ C, the Three deconvoluted peaks of O 1 s in 5% Ag-TiO2 were found
weak acidic contents were 6.22 cm3/g and medium acidic contents (Fig. 4b). The binding energy at 532.4 eV, which belongs to surface
10.68 cm3/g, respectively. Medium-strong acid and medium-strong base oxygen [37], 531.5 eV attributed to oxygen vacancy [38], 529.7 eV
are the primary existing states of 5% Ag-TiO2. Total acid-base content attributed to lattice oxygen (O2− ) in TiO2 or another metal oxide [39],
was listed in Table 2, the acidic content increased significantly from 6.81 respectively. It is worth noting that lattice oxygen accounts for the
to 16.76 cm3/g in 5% Ag-TiO2 as compared to pure TiO2. However, the highest proportion. From the previous work, lattice oxygen is more
alkali content was almost unchanged. stable than adsorbed oxygen and oxygen vacancy. The high ratio of
The element arrangement and crystal structure of 5% Ag-TiO2 lattice oxygen plays a vital role in the subsequent catalytic stability
nanoparticles and pure TiO2 nanoparticles were characterized by X-ray performance.
diffraction (XRD) (Fig. 3). The diffraction of pure TiO2 (the red line), The XPS peaks located around ~464.2 eV and ~458.6 eV ascribed to

Fig. 4. XPS analysis of C 1 s (a), O 1s (b), Ti 2p (c), and Ag 3d core level spectra (d) in fresh 5% Ag-TiO2.

4
T. Xie et al. Molecular Catalysis 525 (2022) 112353

Fig. 5. SEM of 5% Ag-TiO2 (a) and HR-TEM results of 5% Ag-TiO2 (b and c).

the typical binding energies for Ti 2p3/2 and Ti 2p1/2 core levels in of reaction [29]. The XPS results of Ti and O core level spectra did not
anatase TiO2 [40] (Fig. 4c). It can be seen that the addition of Ag did not change significantly.
change the lattice structure of TiO2, and Ti in the sample mainly existed SEM was performed for the morphology and elemental composition
in the form of Ti4+ [41]. As revealed in Fig. 4d, Ag 3d XPS spectrum of analysis (Fig. 5a). Ag-TiO2 showed irregular block structure
5% Ag-TiO2 consists of two bands assigned to the 3d3/2 and 3d5/2 morphology. Meanwhile, Ag, Ti, C and O entities were homogeneously
transitions. Another two different deconvoluted peaks around at 367.8 distributed on the surface of the catalyst. It confirmed the successful
and 373.8 eV belonged to the silver oxide Ag2O, and deconvoluted peaks incorporation of Ag species into TiO2. The spent catalyst SEM charac­
around at 374.4 and 368.9 eV were assigned to the Ag0 metallic [42]. terization was recorded in Fig. S2. After reaction Ag, Ti, C and O entities
From Ti and Ag elements analysis of XPS, Ti4+ and Ag+ existed as TiO2 were also homogeneously distributed, while their size are much smaller
and Ag2O as well a small fraction of Ag0 metal in the prepared than fresh catalyst.
composite. The measurement of microstructure and nanoparticle size of 5% Ag-
Meanwhile, the spent catalyst of C, O, Ti, Ag core level spectra after TiO2 were further studied by HR-TEM (Fig. 5b and c). A uniform quasi-
reaction (before calcination) in XPS were recorded in Fig. S1. Three spherical nanoparticles distribution with the average particle size
peaks deconvoluted peaks associated with C 1s moved negligibly from around ca.11.5 nm was observed in Fig. 5b. The interplanar spacing of
the observation of Fig. S1 (a), whereas, the peak located around C–O 0.308 nm assigned to the anatase TiO2 (101) plane in XRD [43]. Besides,
characteristic peaks (286.4 eV) augmented obviously. It indicated a the Ag2O was detected by the (100) reflection, corresponding to the
small amount of reagent or cokes adsorbed on the surface of the catalyst, spacing d = 0.260 nm. No metallic Ag0 (200) with a spacing of d = 0.208
even after post catalyst washing process. Another obvious change nm and (111) d = 0.234 nm [44] were detected. It indicated the low
focused on Ag (0) position, two deconvoluted peaks around at 374.4 and content of metallic Ag0 in our case, which in accordance with XPS
368.9 eV were assigned to the Ag0 metallic in fresh catalyst (Fig. 4d), results.
whereas, the peaks moved to 375.1 and 369.1 eV in spent one (Fig. S1d).
It suggested that partial Ag2O were reduced to metallic Ag in the process

5
T. Xie et al. Molecular Catalysis 525 (2022) 112353

Fig. 6. (a) Effect of temperature. Reaction condition: 37 mg HMF, 25 mg catalyst, 4 eq. Na2CO3, 30 mL H2O, oxygen pressure 1 MPa, 200 rpm, 16 h, 70–100 ◦ C. (b)
Effect of alkali content. Reaction condition: 37 mg HMF, 25 mg catalyst, 30 mL H2O, oxygen pressure 1 MPa, 80 ◦ C, 200 rpm, 16 h, 0–6 eq. Na2CO3. (c) Effect of
oxygen pressure. Reaction condition: 37 mg HMF, 25 mg catalyst, 30 mL H2O, 80 ◦ C, 0–6 eq. Na2CO3, 200 rpm, 16 h, oxygen pressure 0–1.5 MPa. (d) Effect of
catalyst content. Reaction condition: 37 mg HMF. 30 mL H2O, oxygen pressure 1 MPa, 80 ◦ C, 0–6 eq. Na2CO3, 6–50 mg catalyst, 200 rpm, 16 h.

Fig. 8. Recyclability of regenerate Ag-TiO2. Reaction condition: 37 mg HMF,


Fig. 7. Effect of reaction time. Reaction condition: 37 mg HMF, 25 mg catalyst,
25 mg catalyst, 30 mL H2O, oxygen pressure 1 MPa, 80 ◦ C, 4 eq. Na2CO3, 32 h
30 mL H2O, oxygen pressure 1 MPa, 80 ◦ C, 4 eq. Na2CO3, reaction time ranged
under 200 rpm.
from 2 to 32 h under 200 rpm.

6
T. Xie et al. Molecular Catalysis 525 (2022) 112353

Scheme 1. Pathways for the degradation of lignocellulose and HMF oxidation derivatives.

3.2. Catalytic activity – influence of temperature, base amount, oxygen conversion was found (50%), whereas, HMFCA yield was merely 8%
pressure, catalyst amounts and time under 6 mg Ag-TiO2. It indicated with insufficient Ag active sites, more
HMF degradation happened under basic condition. Further, with the
As depicted in Fig. 6(a), HMF conversion reached ca. 30%, 20% increase of catalyst content, 96% HMFCA selectivity, 72% HMF con­
HMFCA yield and satisfactory HMFCA selectivity (67%) at 70 ◦ C. When version and 70% HMFCA yield obtained with 25 mg catalyst. Contin­
the temperature increased from 70 to 80 ◦ C, the selectivity of HMFCA uous of catalyst amount to 50 mg, HMF conversion, as well as HMFCA
reached the maximum value (98%), HMF conversion (70%) and HMFCA yield reached the maximum (85% and 80%), and the highest HMFCA
yield (68%). Further increment of temperature to 90 and 100 ◦ C, the selectivity (95%). However, compared to 25 mg catalyst performance, in
selectivity of HMFCA decreased to 85% and 57%, it proved 80 ◦ C was the perspective of cost and effect, 25 mg catalyst amount was used for
the optimum temperature despite the conversion of HMF and HMFCA the following optimization. As compared to 25 mg pure TiO2, both HMF
yield reached highest at 90 ◦ C (HMF conversion 87%, HMFCA yield conversion, HMFCA yield and selectivity were lower than 50%. It indi­
74%). This phenomenon related to higher temperature caused the cated the importance of Ag noble metal to the reaction. Meanwhile, the
degradation of HMF, which goes well with the previous work [45]. result hinted the synergistic effect between Ag and TiO2 in HMF selec­
Alkali environment played an essential role in HMF oxidation con­ tive oxidation to HMFCA.
version. As illustrated in Fig. 6(b), when the base amount was 0 eq., the Finally, the effect of reaction time was performed under optimum
selectivity of HMFCA was merely 25% (HMF conversion 20%, HMFCA conditions, as illustrated in Fig. 7. Reaction time in 2 h corresponding to
yield 5%). With the increase of alkali content from 0 to 4 eq., these HMF conversion (6%) and HMFCA yield (5%), respectively. Gradually
values reached the maximum (HMF conversion 70%, HMFCA selectivity increment of the reaction time, the results augment dramatically in 8 h.
98%, HMFCA yield 68%). As the alkali concentration further arrived in 6 Noticeably, HMFCA selectivity reached more than 95%. It proved the
eq., HMF conversion and HMFCA yield and selectivity began to decline. proper reaction time played an important role in HMF conversion. When
Therefore, the optimal reaction concentration was located at 4 eq. the reaction time increased to 32 h, HMF conversion reached 99%, and
Oxygen pressure no doubt can accelerate the reaction process. Fig. 6c HMFCA yield and selectivity all arrived around 96%.
showed the effect of oxygen pressure to the reaction. From the obser­ The stability of catalyst was vital to its application, the recyclability
vation, HMF conversion and HMFCA selectivity achieved 50% without of 5% Ag− TiO2 was tested (Fig. 8). In each successive cycle, the catalyst
oxygen pressure, it proved Ag+ was the active site to this reaction even was centrifuged and washed with water and ethanol three times and
though without oxygen, so-called “silver mirror” reaction. Gradually then prepared by calcination at 350 ◦ C for 3 h. The recovered catalyst
increase the reaction pressure from 0.5 to 1 MPa, the selectivity of had a slight decline in terms of conversion and yield after the third or
HMFCA reached a maximum of 98% (HMF conversion 70%, HMFCA fourth consecutive runs. However, HMF conversion and HMFCA yield
yield 68%). No obvious increment in HMF conversion and HMFCA yield still remained above 90% and 85%, respectively. It indicated that the
and selectivity were found in 1.5 MPa (HMF conversion 80%, HMFCA catalyst has a slight deactivation. From previous work, the negligible
yield 73% with 93% selectivity), this ascribes to the short reaction time deactivation may be related to coke residue deposited on the catalyst
caused no big difference when oxygen source augment. Thus, from the [24].
perspective of cost and effect, 1 MPa was chosen for the following
optimization. 4. Conclusion
Catalyst amount was investigated in Fig. 6d. As exhibited in the
graph, HMFCA selectivity was low (15%), a relative higher HMF TiO2 and 5% Ag-TiO2 were successfully prepared by a simple sol-gel

7
T. Xie et al. Molecular Catalysis 525 (2022) 112353

hydrothermal method and studied for the selective oxidation of HMF to furancarboxylic acid by a new whole-cell biocatalyst Pseudomonas aeruginosa PC-1,
React. Chem. Eng. 5 (2020) 1397–1404, https://doi.org/10.1039/D0RE00018C.
prepare HMFCA. Different reaction parameters were investigated, 99%
[8] M. Munekata, G. Tamura, The selective inhibitors against SV40-transformed cells.
HMF conversion with 96% HMFCA yield, and 97% HMFCA selectivity Antitumor activity of 5-hydroxymethyl-2-furoic acid, Agric. Biol. Chem. 45 (1981)
achieved under optimum conditions. The synthesized catalysts were 2149–2150, https://doi.org/10.1271/bbb1961.45.2149.
characterized by several characterizations. Through these analyses, Ag, [9] A.C. Braisted, J.D. Oslob, W.L. Delano, J. Hyde, R.S. McDowell, N. Waal, C. Yu, M.
R. Arkin, B.C. Raimundo, Discovery of a potent small molecule IL-2 inhibitor
Ti, and O entities were homogeneously distributed, the combination of through fragment assembly, J. Am. Chem. Soc. 125 (2003) 3714–3715, https://doi.
Ag without changing the crystal structure of TiO2, but increases the org/10.1021/ja034247i.
surface area and significantly improved its acidity. TiO2 combined with [10] T. Su, D. Zhao, Y. Wang, H. Lü, R.S. Varma, C. Len, Innovative protocols in the
catalytic oxidation of 5-hydroxymethylfurfural, ChemSusChem 14 (2021)
Ag could adjust the structural network to achieve the maximum porous 266–280, https://doi.org/10.1002/cssc.202002232.
surface area (129.72 m2/g). Ti4+ and Ag+ exist as TiO2 and Ag2O as well [11] D. Zhao, T. Su, Y. Wang, R.S. Varma, C. Len, Recent advances in catalytic oxidation
a small fraction of Ag0 metal in the prepared composite. Catalyst showed of 5-hydroxymethylfurfural, Mol. Catal. 495 (2020), 111133, https://doi.org/
10.1016/j.mcat.2020.111133.
good recyclability until fourth recycling, no apparent deactivation was [12] J.A. Cecilia, L. Machogo, V. Torres-Bujalance, C.P. Jiménez-Gómez, C. García-
found. This work provides a simple and efficient catalyst for the Sancho, R. Moreno-Tost, P. Maireles-Torres, R. Luque, PdO supported on TiO2 for
oxidation of HMF into HMFCA under an oxygen atmosphere, which the oxidative condensation of furfural with ethanol: insights on reactivity and
product selectivity, ACS Sustain. Chem. Eng. 9 (2021) 10100–10112, https://doi.
undoubtedly pave the way for the future development of biomass org/10.1021/acssuschemeng.1c01808.
compound valorization. [13] Y.Y. Gorbanev, S.K. Klitgaard, J.M. Woodley, C.H. Christensen, A. Riisager, Gold-
catalyzed aerobic oxidation of 5-hydroxymethylfurfural in water at ambient
temperature, ChemSusChem 2 (2009) 672–675, https://doi.org/10.1002/
CRediT authorship contribution statement cssc.200900059.
[14] S.E. Davis, L.R. Houk, E.C. Tamargo, A.K. Datye, R.J. Davis, Oxidation of 5-
Tao Xie: Methodology, Writing – original draft. Shitong Yue: hydroxymethylfurfural over supported Pt, Pd and Au catalysts, Catal. Today 160
(2011) 55–60, https://doi.org/10.1016/j.cattod.2010.06.004.
Methodology, Writing – original draft. Ting Su: Formal analysis. Mingqi [15] N.K. Gupta, S. Nishimura, A. Takagaki, K. Ebitani, Hydrotalcite-supported gold-
Song: Methodology, Writing – original draft. Wenjie Xu: Methodology, nanoparticle-catalyzed highly efficient base-free aqueous oxidation of 5-hydroxy­
Writing – original draft. Yaxi Xiao: Methodology, Writing – original methylfurfural into 2,5-furandicarboxylic acid under atmospheric oxygen pressure,
Green Chem. 13 (2011) 824–827, https://doi.org/10.1039/c0gc00911c.
draft. Zhenglong Yang: Conceptualization, Writing – review & editing. [16] T. Pasini, M. Piccinini, M. Blosi, R. Bonelli, S. Albonetti, N. Dimitratos, J.A. Lopez-
Christophe Len: Conceptualization, Writing – review & editing. Sanchez, M. Sankar, Q. He, C.J. Kiely, G.J. Hutchings, F. Cavani, Selective
Deyang Zhao: Methodology, Writing – original draft. oxidation of 5-hydroxymethyl-2-furfural using supported gold–copper
nanoparticles, Green Chem. 13 (2011) 2091–2099, https://doi.org/10.1039/
c1gc15355b.
[17] Z. Miao, Y. Zhang, X. Pan, T. Wu, B. Zhang, J. Li, T. Yi, Z. Zhang, X. Yang, Superior
Declaration of Competing Interest catalytic performance of Ce1− xBixO2− δ solid solution and Au/Ce1− xBixO2− δ for 5-
hydroxymethylfurfural conversion in alkaline aqueous solution, Catal. Sci.
There are no conflicts to declare. Technol. 5 (2015) 1314–1322, https://doi.org/10.1039/C4CY01060D.
[18] F. Wang, Z. Zhang, Cs-substituted tungstophosphate-supported ruthenium
nanoparticles: an effective catalyst for the aerobic oxidation of 5-hydroxymethyl­
Acknowledgments furfural into 5-hydroxymethyl-2-furancarboxylic acid, J. Taiwan Inst. Chem. Eng.
70 (2017) 1–6, https://doi.org/10.1016/j.jtice.2016.10.003.
[19] Z. Zhang, B. Liu, K. Lv, J. Sun, K. Deng, Aerobic oxidation of biomass derived 5-
We are grateful for the funding supported by the National Natural hydroxymethylfurfural into 5-hydroxymethyl-2-furancarboxylic acid catalyzed by
Science Foundation of China (NSFC Grant Nos. 51673090, 52173075 a montmorillonite K-10 clay immobilized molybdenum acetylacetonate complex,
and 22005262). D. Zhao would like to thanks Ludong University for its Green Chem. 16 (2014) 2762–2770, https://doi.org/10.1039/c4gc00062e.
[20] O.R. Schade, K.F. Kalz, D. Neukum, W. Kleist, J.D. Grunwaldt, Supported gold- and
research start-up funding (20210106). silver-based catalysts for the selective aerobic oxidation of 5-(hydroxymethyl)
furfural to 2,5-furandicarboxylic acid and 5-hydroxymethyl-2-furancarboxylic
acid, Green Chem. 20 (2018) 3530–3541, https://doi.org/10.1039/C8GC01340C.
Supplementary materials
[21] O.R. Schade, F. Stein, S. Reichenberger, A. Gaur, E. Saraҫi, S. Barcikowski,
J. Grunwaldt, Selective aerobic oxidation of 5-(Hydroxymethyl)furfural over
Supplementary material associated with this article can be found, in heterogeneous silver-gold nanoparticle catalysts, Adv. Synth. Catal. 362 (2020)
the online version, at doi:10.1016/j.mcat.2022.112353. 5681–5696, https://doi.org/10.1002/adsc.202001003.
[22] M. Liu, H. Wang, H. Zeng, C.J. Li, Silver(I) as a widely applicable, homogeneous
catalyst for aerobic oxidation of aldehydes toward carboxylic acids in
References water—“silver mirror”: from stoichiometric to catalytic, Sci. Adv. 1 (2015),
e1500020, https://doi.org/10.1126/sciadv.1500020.
[23] D. Zhao, D. Rodriguez-Padron, R. Luque, C. Len, Insights into the selective
[1] A.J.J.E. Eerhart, A.P.C. Faaij, M.K. Patel, Replacing fossil based PET with biobased
oxidation of 5-hydroxymethylfurfural to 5-hydroxymethyl-2-furancarboxylic acid
PEF; process analysis, energy and GHG balance, Energy Environ. Sci. 5 (2012)
using silver oxide, ACS Sustain. Chem. Eng. 8 (2020) 8486–8495, https://doi.org/
6407–6422, https://doi.org/10.1039/c2ee02480b.
10.1021/acssuschemeng.9b07170.
[2] C. Chen, L. Wang, B. Zhu, Z. Zhou, S.I. El-Hout, J. Yang, J. Zhang, 2,5-
[24] T. Su, Q. Liu, H. Lü, F.Ali Alasmary, D. Zhao, C. Len, Selective oxidation of 5-
Furandicarboxylic acid production via catalytic oxidation of 5-hydroxymethylfur­
hydroxymethylfurfural to 5-hydroxymethyl-2-furancarboxylic acid using silver
fural: catalysts, processes and reaction mechanism, J. Energy Chem. 54 (2021)
oxide supported on calcium carbonate, Mol. Catal. 502 (2021), 111374, https://
528–554, https://doi.org/10.1016/j.jechem.2020.05.068.
doi.org/10.1016/j.mcat.2020.111374.
[3] C. Xu, E. Paone, D. Rodríguez-Padrón, R. Luque, F. Mauriello, Recent catalytic
[25] C.A. Celaya, R. Oukhrib, M.A. El Had, Y. Abdellaoui, H.A. Oualid, H. Bourzi,
routes for the preparation and the upgrading of biomass derived furfural and 5-
R. Chahboun, D. Zhao, S.M. Osman, V.S. Parmar, C. Len, Density functional theory
hydroxymethylfurfural, Chem. Soc. Rev. 49 (2020) 4273–4306, https://doi.org/
study of the selective oxidation of 5-hydroxymethylfurfural (HMF) to 5-hydroxy­
10.1039/D0CS00041H.
methyl-2-furancarboxylic acid (HMFCA) on the Silver oxide surface (001), Mol.
[4] C.P. Jiménez-Gómez, J.A. Cecilia, C. García-Sancho, R. Moreno-Tost, P. Maireles-
Catal. 519 (2022), 112117, https://doi.org/10.1016/j.mcat.2022.112117.
Torres, Selective production of furan from gas-phase furfural decarbonylation on
[26] Y. Yue, D. Yuchi, P. Guan, J. Xu, L. Guo, J. Liu, Atomic scale observation of oxygen
Ni-MgO catalysts, ACS Sustain. Chem. Eng. 7 (2019) 7676–7685, https://doi.org/
delivery during silver–oxygen nanoparticle catalysed oxidation of carbon
10.1021/acssuschemeng.8b06155.
nanotubes, Nat. Commun. 7 (2016) 12251, https://doi.org/10.1038/
[5] R. Maderuelo-Solera, S. Richter, C.P. Jiménez-Gómez, C. García-Sancho, F.
ncomms12251.
J. García-Mateos, J.M. Rosas, R. Moreno-Tost, J.A. Cecilia, P. Maireles-Torres,
[27] J. An, G. Sun, H. Xia, Aerobic oxidation of 5-hydroxymethylfurfural to high-yield 5-
Porous SiO2 nanospheres modified with ZrO2 and their use in one-pot catalytic
hydroxymethyl-2-furancarboxylic acid by poly(vinylpyrrolidone)-capped Ag
processes to obtain value-added chemicals from furfural, Ind. Eng. Chem. Res. 60
nanoparticle catalysts, ACS Sustain. Chem. Eng. 7 (2019) 6696–6706, https://doi.
(2021) 18791–18805, https://doi.org/10.1021/acs.iecr.1c02848.
org/10.1021/acssuschemeng.8b05916.
[6] B.T. Olea, I.F. Nuñez, C.G. Sancho, J.A. Cecilia, R.M. Tost, P.M. Torres, Production
[28] R. Behling, S. Valange, G. Chatel, Heterogeneous catalytic oxidation for lignin
of biofuels by 5-hydroxymethylfurfural etherification using ion-exchange resins as
valorization into valuable chemicals: what results? What limitations? What trends?
solid acid catalysts, Chem. Proc. 2 (2020) 34–39, https://doi.org/10.3390/
Green Chem. 18 (2016) 1839–1854, https://doi.org/10.1039/C5GC03061G.
ECCS2020-07587.
[29] N. Wei, H. Cui, Q. Song, L. Zhang, X. Song, K. Wang, Y. Zhang, J. Li, J. Wen,
[7] X. Pan, S. Wu, D. Yao, L. Liu, L. Zhang, Z. Yao, Y. Pan, S. Chang, B. Li, Efficient
J. Tian, Ag2O nanoparticle/TiO2 nanobelt heterostructures with remarkable photo-
biotransformation of 5-hydroxymethylfurfural to 5-hydroxymethyl-2-

8
T. Xie et al. Molecular Catalysis 525 (2022) 112353

response and photocatalytic properties under UV, visible and near-infrared Appl. Catal. B Environ. 250 (2019) 313–324, https://doi.org/10.1016/j.
irradiation, Appl. Catal. B Environ. 198 (2016) 83–90, https://doi.org/10.1016/j. apcatb.2019.03.055.
apcatb.2016.05.040. [38] L. Zhang, Q. Liang, P. Yang, Y. Huang, Y. Liu, H. Yang, J. Yan, ZIF-8 derived ZnO/
[30] J.G. Kang, Y. Sohn, Interfacial nature of Ag nanoparticles supported on TiO2 Zn6Al2 O9/Al2O3 nanocomposite with excellent photocatalytic performance under
photocatalysts, J. Mater. Sci. 47 (2012) 824–832, https://doi.org/10.1007/ simulated sunlight irradiation, New J. Chem. 43 (2019) 2990–2999, https://doi.
s10853-011-5860-6. org/10.1039/C8NJ04798G.
[31] T. Su, Y. Yang, Y. Shi, X. Zhang, Y. Jiang, R. Fan, W. Cao, 40% enhanced [39] Y. Wang, Y. Zhang, H. Lu, Y. Chen, Z. Liu, S. Su, Y. Xue, J. Yao, H. Zeng, Novel N-
photocurrent of dye sensitized solar cells using lotus-shaped H 2 -treated anatase doped ZrO2 with enhanced visible-light photocatalytic activity for hydrogen
TiO2 with {001} dominated facets, Chem. Eng. J. 316 (2017) 534–543, https://doi. production and degradation of organic dyes, RSC Adv. 8 (2018) 6752–6758,
org/10.1016/j.cej.2017.01.107. https://doi.org/10.1039/C7RA12938F.
[32] T. Su, Y. Yang, Y. Na, R. Fan, L. Li, L. Wei, B. Yang, W. Cao, An insight into the role [40] J.S. Lee, K.H. You, C.B. Park, Highly photoactive, low bandgap TiO2 nanoparticles
of oxygen vacancy in hydrogenated TiO2 nanocrystals in the performance of dye- wrapped by graphene (Adv. Mater. 8/2012), Adv. Mater. 24 (2012) 1133, https://
sensitized solar cells, ACS Appl. Mater. Interfaces 7 (2015) 3754–3763, https://doi. doi.org/10.1002/adma.201290038, 1133.
org/10.1021/am5085447. [41] M.R. Uddin, M.R. Khan, M.W. Rahman, A. Yousuf, C.K. Cheng, Photocatalytic
[33] K.S.W. Sing, Reporting physisorption data for gas/solid systems with special reduction of CO2 into methanol over CuFe2O4/TiO2 under visible light irradiation,
reference to the determination of surface area and porosity (Recommendations React. Kinet. Mech. Catal. 116 (2015) 589–604, https://doi.org/10.1007/s11144-
1984), Pure Appl. Chem. 57 (1985) 603–619, https://doi.org/10.1351/ 015-0911-7.
pac198557040603. [42] T. Bansala, S. Mukhopadhyay, M. Joshi, R. Doong, M. Chaudhary, Synthesis and
[34] K.A. Jacob, P.M. Peter, P.E. Jose, C.J. Balakrishnan, V.J. Thomas, A simple method shielding properties of PVP-stabilized-AgNPs-based graphene nanohybrid in the Ku
for the synthesis of anatase-rutile mixed phase TiO2 using a convenient precursor band, Synth. Met. 221 (2016) 86–94, https://doi.org/10.1016/j.
and higher visible-light photocatalytic activity of Co–doped TiO2, Mater. Today synthmet.2016.07.034.
Proc. 49 (2022) 1408–1417, https://doi.org/10.1016/j.matpr.2021.07.104. [43] E. Assayehegn, A. Solaiappan, Y. Chebude, E. Alemayehu, Fabrication of tunable
[35] L. Nossova, G. Caravaggio, M. Couillard, S. Ntais, Effect of preparation method on anatase/rutile heterojunction N/TiO2 nanophotocatalyst for enhanced visible light
the performance of silver-zirconia catalysts for soot oxidation in diesel engine degradation activity, Appl. Surf. Sci. 515 (2020), 145966, https://doi.org/
exhaust, Appl. Catal. B Environ. 225 (2018) 538–549, https://doi.org/10.1016/j. 10.1016/j.apsusc.2020.145966.
apcatb.2017.11.070. [44] E. Albiter, M.A. Valenzuela, S. Alfaro, G. Valverde-Aguilar, F.M. Martínez-Pallares,
[36] M.K. Rabchinskii, S.D. Saveliev, D.Yu. Stolyarova, M. Brzhezinskaya, D. Photocatalytic deposition of Ag nanoparticles on TiO2: metal precursor effect on
A. Kirilenko, M.V. Baidakova, S.A. Ryzhkov, V.V. Shnitov, V.V. Sysoev, P. the structural and photoactivity properties, J. Saudi Chem. Soc. 19 (2015)
N. Brunkov, Modulating nitrogen species via N-doping and post annealing of 563–573, https://doi.org/10.1016/j.jscs.2015.05.009.
graphene derivatives: XPS and XAS examination, Carbon 182 (2021) 593–604, [45] D. Zhao, D. Rodriguez-Padron, K.S. Triantafyllidis, Y. Wang, R. Luque, C. Len,
https://doi.org/10.1016/j.carbon.2021.06.057. Microwave-assisted oxidation of hydroxymethyl furfural to added-value
[37] G. Zhang, D. Chen, N. Li, Q. Xu, H. Li, J. He, J. Lu, Fabrication of Bi2MoO6/ZnO compounds over a ruthenium-based catalyst, ACS Sustain. Chem. Eng. 8 (2020)
hierarchical heterostructures with enhanced visible-light photocatalytic activity, 3091–3102, https://doi.org/10.1021/acssuschemeng.9b05656.

You might also like