You are on page 1of 8

Catalysis Today 379 (2021) 62–69

Contents lists available at ScienceDirect

Catalysis Today
journal homepage: www.elsevier.com/locate/cattod

Tin, niobium and tin-niobium oxides obtained by the Pechini method using
glycerol as a polyol: Synthesis, characterization and use as a catalyst in
fructose conversion
Thatiane V. dos Santos , Dhara B.A. Pryston , Geovânia C. Assis , Mario R. Meneghetti , Simoni M.
P. Meneghetti *
Group of Catalysis and Chemical Reactivity, Institute of Chemistry and Biotechnology, Federal University of Alagoas, Av. Lourival de Melo Mota, s/ nº, Maceió, AL,
57072-970, Brazil

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

Keywords: The selective and efficient conversion of fructose into chemicals has been extensively explored, especially from
Biomass the perspective of a more sustainable industry based on renewable inputs together with a more environmentally
Fructose friendly process. In this study, mixed metal oxides based on tin and niobium were synthesized by the Pechini
Pechini
method, replacing ethylene glycol with glycerol, a renewable chemical. This replacement allows the use of pure
Catalyst
or mixed oxides exhibiting high specific surface areas but did not lead to significant differences in the amount
Mixed oxide
Tin and nature of the acid sites present in each case. Furthermore, the mixed oxides had increased numbers of Lewis
Niobium acid sites, while the number of Bronsted acid sites decreased. Catalytic tests performed at 150 ◦ C demonstrated
that the use of these materials leads to conversion results higher than those observed for pure oxides (SnO2 (EG)
= 48.3 %, Nb2O5 (EG) = 51.6 % and SnO2 (G) = 56.8 %, Nb2O5 (G) = 59.6 %) because 78.5 and 75.7 %
consumption of fructose for SnNb (EG) and SnNb (G) was attained at 2 h, respectively. In addition, the catalysts
show promising 5-hydroxymethylfurfural (5-HMF) selectivity, and the presence of lactic acid was also detected.
In addition, recycling experiments were performed, and stability was observed for 4 cycles without structural
change in the catalysts.

1. Introduction may be followed to the detriment of others.


An important example of a chemical that can be produced by a
With increasing interest in sustainable sources of energy and chem­ biorefinary is 5-hydroxymethylfurfural (5-HMF), which can be trans­
icals, there is a strong focus on the use of biomass resources. A focus is formed into a great range of products of interest, such as 2,5-furandicar­
lignocellulosic biomass, a renewable, low cost and highly available boxylic acid, 2,5-dimethylfuran, γ-valerolactone and levulinic acid,
input that is basically composed of carbohydrate polymers (cellulose which can be used in the polymer, surfactant, pharmaceutical and fuel
and hemicellulose) and an aromatic polymer (lignin) [1,2]. fields [8].
In particular, sugars derived from biomass, such as glucose and Thus, the search for active and selective catalysts able to transform
fructose, can be employed as raw materials to produce biofuels and sugars into products of interest is increasing. In the literature, the use of
chemicals, which can supplement or substitute some petrochemical inorganic acid catalysts [9], ionic liquids [10,11], Sn(IV) complexes [12,
derivatives [3,4]. In fructose conversion in aqueous medium, different 13], silica-included heteropolyacids [2], and inorganic and organic
reaction pathways may occur, such as (i) fructose isomerization to potassium salts [14], among others, has been reported for research on
glucose, (ii) fructose dehydration to 5-hydroxymethylfurfural (5-HMF), biomass conversion. The use of heterogeneous catalysts has many ad­
(iii) rehydration of 5-HMF to levulinic and formic acid, and (iv) vantages over homogeneous catalysts, mainly due to the possibility of
retro-aldol condensation of fructose into lactic acid, acetic acid, dihy­ recovery and reuse, reducing the costs of the final products. As an
droxyacetone, etc. [5–7]. Depending on the reaction conditions, espe­ example, recent studies describe the use of modified beta zeolites in the
cially the nature of the catalysts employed, one of these reaction paths conversion of lignocellulose and carbohydrates to furfural, and different

* Corresponding author.
E-mail address: simoni.plentz@iqb.ufal.br (S.M.P. Meneghetti).

https://doi.org/10.1016/j.cattod.2020.07.002
Received 9 November 2019; Received in revised form 1 June 2020; Accepted 2 July 2020
Available online 23 July 2020
0920-5861/© 2020 Elsevier B.V. All rights reserved.
T.V. dos Santos et al. Catalysis Today 379 (2021) 62–69

yields were obtained according to the nature of the investigated catalytic %, Sigma-Aldrich), ammonium hydroxide (Sigma-Aldrich), nitric acid
system and reaction conditions [15]. (>65.0 %, Sigma-Aldrich), glycerol (> 99.0 %, Sigma-Aldrich), ethylene
However, few examples are available in the literature concerning the glycol (> 99.0 %, Sigma-Aldrich) and fructose (> 99.0 %, Sigma-
use of mixed oxides and their application to biorefinery routes. The use Aldrich) were obtained commercially and used as received.
of CeO2-Nb2O5, which contains mainly Bronsted acid sites, leads to the
conversion of fructose and selectivity for HMF as the amount of niobium 2.2. Synthesis and characterization of catalysts
in the material increases [16]. Another example is based on the use of
mesoporous NbxW(8-x) mixed oxides for the conversion of glucose to The oxides were synthesized using the polymeric precursor method
fructose, mannose and 5-HMF. The best results were obtained using [21] following three consecutive steps: (i) metal citrate preparation: tin
Nb4W4 with a 36.1 % glucose conversion and 18.8 % yield of HMF [17]. or niobium chloride and citric acid solutions (molar ratio metal chloride:
Zr-(W,Al) mixed oxides were used in D-xylose conversion (98 %) to citric acid = 1:3) were mixed and left stirring until reaching 70 ◦ C. After,
produce furfural with yields of 51 % [18]. Additionally, mixed oxides the solution was immersed in an ice bath to initiate the precipitation.
based on SnW and SnMo were investigated in glycerol acetalization, and Subsequently, the solution was returned to the stirring system, and
a selectivity of 96 % could be achieved for the SnM catalyst due to the ammonium hydroxide was added gradually for 4 h until complete pre­
acidic character of this material [19]. Sn-modified Nb2O5, obtained by cipitation was achieved. Then, the metal citrate was filtered, washed and
coprecipitation, was employed for the conversion of triose sugar and dried at 120 ◦ C; (ii) Preparation of the polymeric resin: metal citrate was
yields of 98.7 % of lactic acid was obtained at 160 ◦ C and 4 h in aqueous dissolved in distilled water and nitric acid. Then, polyol (glycerol or
medium [20]. Recently, dos Santos et al. evaluated a series of SnO2 and ethylene glycol) was added at a 60:40 M ratio, considering the amount of
MoO3 oxides with different molybdenum proportions for fructose con­ citric acid used. The solution was heated to 70 ◦ C with stirring to pro­
version. Among the oxides employed, better conversions and significant mote the polyesterification reaction. After elimination of nitrous oxides
selectivity for lactic acid and its formation intermediates were obtained and water, a white resin was obtained. (iii) Heat treatment: finally, the
with the mixed oxide containing 25 % of MoO3 at 150 ◦ C, with less solid was ground and calcined at a temperature of 500 ◦ C for 4 h.
formation of insoluble materials (humines) [7]. In addition, the These systems are named SnO2 (EG), Nb2O5 (EG) and SnNb (EG) for
robustness of mixed oxides and their potential reuse were observed [7]. systems that use ethylene glycol, while for glycerol, the materials will be
In all cases, the use of mixed oxides offers various Lewis and Bronsted named SnO2 (G), Nb2O5 (G) and SnNb (G), respectively. Additionally, a
acids, which play an important role in the conversion and selectivity of sample of mixed oxide was synthesized with glycerol using a 1:1 mass
the catalytic systems. ratio, considering the amount of citric acid used (named as SnNb (G)**).
There are several mixed oxide synthesis routes to obtain the best All materials were characterized by various techniques (see supple­
structure-property relationships. One noteworthy route is the Pechini mentary information).
method, which is based on the use of chelating ligands to form a ho­
mogeneous solution of metal/citrate complexes [21]. Afterwards, this 2.3. Conversion of fructose
solution is reacted with a polyol, usually ethylene glycol (diol), to
convert the mixture into a covalent polymer network to entrap the All experiments were performed in vials of 4 mL under magnetic
metal. This method promotes a homogeneous dispersion of metals stirring and heating at several reaction times. The fructose solution was
throughout the polymer network due to its controlled growth. Subse­ 0.016 g of fructose in 2 mL of deionized water, and for some experi­
quently, the slow decomposition of the polymeric matrix allows greater ments, catalysts (1.5 × 10− 3 g) were used. Conversion, yield and
control over the material formation [22,23]. Changes in the nature of selectivity were calculated from the results of quantification by HPLC
the polyol or acid used in the synthesis may lead to significant changes in (see supplementary information). The reuse tests of SnNb (EG) or SnNb
the properties of the materials obtained. Rudisill et al. investigated the (G) were carried out at 150 ◦ C for 1 h under the same reaction condi­
replacement of citric acid and ethylene glycol by malic acid and glyc­ tions. Subsequently, the catalyst was removed from the reaction medium
erol, respectively, to synthesize polymethylmethacrylate (PMMA)-tem­ by centrifugation and calcined at 550 ◦ C for 4 h for subsequent reuse (see
plated systems, and changes in the material morphology were observed supplementary information).
[22,24].
Glycerol, a renewable raw material, is a coproduct of biodiesel 3. Results and discussion
obtention and is formed in approximately 10 % in the transesterification
process. There is currently great interest in the use of this material for As already mentioned, different catalytic systems based on tin and
the production of high-value-added products of industrial interest, such niobium oxides were obtained via the Pechini method using glycerol or
as acetins, acrolein, 1,3-propanediol, monoacylglycerols (MAGs) and ethylene glycol as polyols in the same synthesis protocol.
diacylglycerols (DAGs) [25,26].
In this work, the use of glycerol to replace ethylene glycol in the 3.1. Characterization of catalysts
synthesis with SnO2, Nb2O5 and a mixed oxide was investigated, and the
materials were evaluated for fructose conversion. To the best of our After the Pechini synthesis, the samples were calcined at 500 ◦ C for 4
knowledge, this is the first time that mixed tin and niobium oxides have h to eliminate the polymeric matrix and to form the materials. After this
been synthesized using glycerol, which can be a renewable source of procedure, the thermal behaviors of the samples were evaluated
polyol, as a polyol in the Pechini synthesis and applied to this bio­ (Fig. S1, supplementary information), and only small mass losses
refinery route. Firstly, the main goal of this study was to assess whether (0.8–3.6 %) were observed in the range of 30–100 ◦ C due to water
it is really possible to exchange ethylene glycol for glycerol and, sec­ molecule elimination, indicating total combustion of the organic mate­
ondly, to evaluate whether, for example, textural properties, acidity, rial. The absence of remains from organic matrix is confirmed by the
among others, may be influenced by this change. spectra obtained in the mid-infrared region (Fig. S2, supplementary
information) where no absorption bands, in the region of 2850 to 3000
2. Materials and methods cm− 1 due to CH– stretching, are detected.
Further, in the FTIR spectrum of SnO2 (EG and G) was observed the
2.1. Materials presence of a broadband in the range of 497 to 662 cm− 1 which is related
to the stretching of the O-Sn-O and Sn-O bonds, respectively [27,28]. In
Tin(IV) chloride (99.995 % trace metals basis, Sigma-Aldrich), the case of Nb2O5 (EG and G), in the range of 410 to 931 cm− 1 is detected
niobium(V) chloride (>99.995 %, Sigma-Aldrich), citric acid (> 99.0 a broadband that is attributed to the symmetrical and asymmetric

63
T.V. dos Santos et al. Catalysis Today 379 (2021) 62–69

Fig. 1. Powder X-ray diffraction patterns of SnO2 (G), Nb2O5 (G) and SnNb (G) (A) and SnO2 (EG), Nb2O5 (EG) and SnNb (EG) (B).

explained by the difference in the nature of the polyol (number of hy­


Table 1
droxyl groups) and the consequent characteristics of the polymeric
Textural and structural properties of the materials.
matrix and the formed material (see the DRX profile of SnNb (G)** in the
Sample SBETa Vb DBJHc Fig. S3).
(m2 g− 1) (cm3 g− 1) (nm)
Fig. S4 (supplementary information) shows the nitrogen adsorption-
SnO2 (EG) 30 0.116 96.0 desorption isotherms obtained at 77 K and the pore diameter distribu­
SnO2 (G) 70 0.110 51.0 tions of the oxides investigated in this work, and all materials exhibit
Nb2O5 (EG) 69 0.124 36.0
Nb2O5 (G) 85 0.129 50.8
type IV isotherms with H1 hysteresis, characteristics of mesoporous
SnNb (EG) 156 0.192 40.5 materials; type IV isotherms with a substantial hysteresis loop indicate
SnNb (G) 167 0.184 40.4 the mesoporous structure of the samples according to the IUPAC clas­
SnNb (G)** 165 0.150 36.4 sification [34,35].
The specific surface areas, total pore volumes, and pore diameters were calcu­ Table 1 presents the textural and structural properties of the solids.
lated by the BET and BJH equations. For all materials, as expected, in their synthesis, the use of glycerol
a
SBET, BET specific surface area. instead of ethylene glycol leads to an increase in the specific surface
b
V, pore volume. area, indicating the influence of the polymeric matrix characteristics on
c
DBJH, pore diameter. the properties of the materials [22,24].
This effect is more pronounced in the case of pure SnO2 because areas
stretching of the Nb-Nb-O and O-Nb-O bonds, respectively [29,30]. For of 30.0 and 69.6 m2 g− 1 are obtained for SnO2 (EG) and SnO2 (G),
both mixed oxides, a broad band in the range of 410 to 931 cm− 1 was respectively. In comparison to commercial SnO2, which displays a spe­
also observed, related to the symmetrical and asymmetric stretching of cific surface area of approximately 12.0 m2 g− 1 [7], the surface area is
the metal bonds with oxygen (O–MO and M––O). For all materials, the much higher. For the mixed oxides, SnNb (EG), SnNb (G) and SnNb (G)
presence of an absorption band at 1620 cm− 1 and 3412 cm− 1 refers to **, higher areas than the pure oxides were obtained, suggesting that the
the angular deformation of the water adsorbed on the material surfaces mutual interaction between metal oxides strongly inhibits their indi­
and to the hydroxyl stretching, respectively [31,32]. vidual crystallization and particle growth, resulting in specific surface
In the case of mixed oxides, the niobium amounts were quantified by areas larger than those of the pure metal oxides [36]. Therefore, the
ICP-OES and, as expected, 19.0 and 16.9 % of Nb were determined for increased surface area of mixed oxides can be expected to be one of the
SnNb (EG) and SnNb (G), respectively. vital reasons for achieving better catalytic activity in biomass conver­
Fig. 1 presents the diffractograms of the materials. SnO2 (EG) and sion [37].
SnO2 (G) exhibit a tetragonal phase (rutile), which is confirmed by the The results from the Raman spectroscopy of the oxides are presented
crystalline planes assigned to the reflection lines (110), (101), (200), in Fig. 2, and the positions and shapes of the signals observed are
(211), (220), (002) (310), (112), (301), (202) and (321) (JCPDS 41- practically unchanged when ethylene glycol is replaced by glycerol in
1445). The DRX patterns for Nb2O5 (EG) and Nb2O5 (G) show the the synthesis. The Raman spectra of SnO2 (EG) and SnO2 (G) (Fig. 2A)
reflection lines (001), (180), (181), (002), (321), (182), (202), (2160), mainly exhibit three bands at 474, 630 and 770 cm− 1, which are
(2161) and (382), indicating mainly the presence of an orthorhombic attributed to the vibration modes Eg, A1g and B2g, respectively. The
phase (T Nb2O5 - JCPDS 30-0873). However, it is not possible to discard presence of these bands corroborates the tetragonal structure of SnO2,
the presence, in a lower content, of a pseudohexagonal phase with which is in agreement with the DRX results. In addition to the three
reflection lines at (001), (100), (101), (002), (110) and (102) (TT Nb2O5 strong signals, there is (i) another weak signal at approximately 694
- JCPDS 280317). These phases exhibit relatively similar X-ray diffrac­ cm− 1, which is attributed to the IR-active A2u LO mode (the mode of
tion patterns, and some studies suggest that TT-Nb2O5 is a less crystal­ longitudinal optical phonons), and (ii) a signal at 578 cm− 1 related to
line form of T-Nb2O5 stabilized by impurities or oxygen vacancies [33]. surface phonon modes [38]. Thus, the Raman results provide evidence
For SnNb (EG), SnNb (G) and SnNb (G)**, it is possible to observe the that the surfaces of both samples contain oxygen vacancies [35,37]. The
preponderant presence of the tetragonal phase, and the signals are Nb2O5 (EG) and Nb2O5 (G) spectra are shown in Fig. 2B and reveal the
broader and less intense in comparison with the pure oxides, suggesting weak Raman band located at 990 cm− 1 due to the vibration modes A1g
that the difference in the ionic radii due to the substitution of Sn4+ by (ν1) of NbO6 octahedron in Nb2O5 [37,39]. Signals that appear between
Nb5+ induces a distortion in the lattice constants of the SnO2 unit cell. 200 and 300 cm− 1 are characteristic of the flexion modes of Nb–O–Nb
Nevertheless, the DRX profiles for materials obtained using glycerol bonds [40]. The Raman band at approximately ~700 cm− 1 can be
during synthesis showed broader and less intense signals, which can be assigned to the orthorhombic Nb2O5 TO mode [39].

64
T.V. dos Santos et al. Catalysis Today 379 (2021) 62–69

Fig. 2. Raman spectra of SnO2 (G) and SnO2 (EG) (A), Nb2O5 (G) and Nb2O5 (EG) (B) and SnNb (G) and SnNb (EG) (C).

Fig. 3. Absorption spectra (UV–vis DRS) of SnO2 (G), Nb2O5 (G) and SnNb (G) (A) and SnO2 (EG), Nb2O5 (EG) and SnNb (EG) (B).

For the mixed oxides SnNb EG and SnNb G (Fig. 2C), a signal at (present in Nb2O5 G or EG), and (iii) the presence of an absorption band
800− 900 cm− 1 is observed due to symmetrical stretching of Nb = O between 220 and 250 nm, which is related to the overlap of tetrahedral
species present in distorted NbO6 octahedra [41,42]. The signals at 450 Sn species and the strong interaction with octahedral niobium species
cm− 1 and 632 cm− 1 can be attributed to the SnO2 Eg mode and sym­ [43].
metric A1g mode associated with the elongation of the Sn-O bond, To evaluate the nature of the acid sites present in these materials,
respectively [42]. The presence of the Eg mode is a strong indication of they were analyzed by infrared spectroscopy using pyridine as a probe
oxygen vacancies. Nevertheless, the signal at 256 cm− 1 can be related to molecule. The absorption bands at approximately 1458, 1507 and 1633
the formation of pure tin oxide nanoparticles [20]. cm− 1 are related to the presence of Lewis acid sites, and those at 1559
By UV–vis DRS (Fig. 3), it is evident that the spectra of both SnNb EG and 1652 cm − 1 are related to Bronsted acid sites [44–48]. At 1483
and SnNb G have different profiles than the pure oxides, proving the cm− 1, the superposition of the Lewis and Bronsted sites is observed [47,
formation of materials exhibiting other structures due to the (i) 48]. It is important to mention that in the case of Nb2O5 (G and EG) is
displacement of the band with a maximum at 300 nm (present in SnO2 G observed a large absorption band, possibly related to the presence to
or EG), demonstrating the absence of bulk SnO2, (ii) the disappearance Lewis sites associated to diverse NbOx species present at material sur­
of the absorption for corner-sharing NbO6 octahedra at ~ 350 nm face, as Nb = O structure unit of tetrahedral mono-oxo NbO4 or

65
T.V. dos Santos et al. Catalysis Today 379 (2021) 62–69

Fig. 4. Number of Lewis (A) and Bronsted acid sites (B) present in the catalytic systems at different temperatures.

acid sites with weak to moderate strength. In terms of the number of


Table 2
sites, SnO2 (G) and SnO2 (EG) present lower values than Nb2O5 (G) and
Lewis / Brönsted acid sites ratio (IL/IB) at different temperatures obtained from
Nb2O5 (EG). The highlight is that the mixed oxides had increases in the
the infrared spectra of pyridine adsorbed on catalysts.
number of Lewis acid sites, while the number of Bronsted acid sites
IL/IB
T decreased. This can best be seen when analyzing IL/IB, where a signifi­
(ºC) SnO2 SnO2 Nb2O5 Nb2O5 SnNb SnNb cant increase is observed for all temperatures. The modulation of these
(EG) (G) (EG) (G) (EG) (G) characteristics in mixed oxides is fundamental for making this system
25 0.7 0.7 0.9 1.0 1.4 1.4 more versatile for application in various biorefinery platforms [20,47].
100 0.8 0.8 0.9 0.8 4.4 4.7
200 1.2 1.2 1.3 2.4 6.0 5.8
3.2. Catalytic tests
oligomeric niobia [49].
SnO2 (EG), Nb2O5 (EG), SnNb (EG), SnO2 (G), Nb2O5 (G) and SnNb
To investigate the acid strength of the acid sites, the thermal stability
(G) were investigated as catalysts for the conversion of fructose in
of the pyridine-acid sites was followed by the acquisition of infrared
aqueous medium at 150 ◦ C and several reaction times (see Fig. 5).
spectra at 100, 200 and 300 ◦ C. When pyridine is chemisorbed at Lewis
First, it is important to mention that in previous works on fructose
or Bronsted acid sites, this interaction can be considered weak (lower
conversion using commercial SnO2 as a catalyst, only yields analogous
than 200 ◦ C), moderate (200–300 ◦ C) or strong (above 300 ◦ C) ac­
to those observed for the reaction run without a catalyst were attained
cording to the temperature required to break the interaction [49,50].
under the same reaction conditions [7]. However, analyzing the results
Furthermore, by Equation 1 (supplementary information), it is possible
obtained employing SnO2 (G) and SnO2 (EG), it is possible to observe
to quantify the number of acid sites at different temperatures (Fig. 4,
significant values of conversion, almost close to those obtained for the
Table 2 and Fig. S5 - supplementary information). According to the
Nb2O5 (G) and Nb2O5 (EG) systems. Furthermore, for the pure oxides
obtained results, it is possible to confirm that all systems present weak to
synthesized using glycerol as a polyol, higher conversions are observed
moderate acid sites since no signals were detected at 300 ◦ C (Fig. 4).
compared with those obtained with ethylene glycol. These trends
In general, the replacement of ethylene glycol for glycerol in the
confirm the importance of obtaining materials with a high surface area,
oxide synthesis process did not lead to significant differences in the
obviously increasing the presence of suitable catalytic sites.
amount and nature of the acid sites present, and all materials exhibit
Concerning the mixed oxides SnNb (EG) and SnNb (G), the

Fig. 5. Fructose conversion at 150 ◦ C using SnO2 (G), Nb2O5 (G) and SnNb (G) (A) and SnO2 (EG), Nb2O5 (EG) and SnNb (EG) (B).

66
T.V. dos Santos et al. Catalysis Today 379 (2021) 62–69

3
Fig. 6. Selectivity for soluble products identified in the fructose conversion at 150 ◦ C without catalyst and with 1.5 × 10− g of catalyst.

67
T.V. dos Santos et al. Catalysis Today 379 (2021) 62–69

3
Fig. 7. Fructose conversions at 120, 150 and 170 ◦ C without or with 1.5 × 10− g of SnNb (G) and SnNb (EG).

conversion results are higher than those observed, for example, for pure increase in the total number of Lewis or Bronsted acid sites in compar­
oxides at 2 h (SnO2 (EG) = 48.3 %, Nb2O5 (EG) = 51.6 % and SnO2 (G) = ison with the individual oxides (Fig. 4 and Table 2). Evaluating these
56.8 %, Nb2O5 (G) = 59.6 %), since 78.5 and 75.7 % consumption of results, pure Nb2O5 (G or EG) exhibits comparable numbers of Lewis and
fructose for SnNb (EG) and SnNb (G) was attained at 2 h, respectively Bronsted acid sites, and the mixed oxides had a significant increase in
(Fig. 5). the number of Lewis sites, which justifies the higher conversion
Regarding the products formed from fructose conversion, glucose, 5- observed for SnNb (G or EG). However, the presence of Bronsted acid
HMF, levulinic acid, formic acid, glyceraldehyde, dihydroxyacetone, sites, even fewer than those observed for pure Nb2O5 (G or EG), gua­
pyruvaldehyde, lactic acid and acetic acid were detected (Table T1 and rantees the same tendency in terms of selectivity.
T2, supplementary information). Fig. 6 shows the selectivity at 150 ◦ C During these reactions, insoluble products (polymeric materials,
for the catalysts tested. All systems exhibit selectivity for 5-HMF (60–80 such as humins) are formed, principally by 5-HMF decomposition, and
%) and, to a lesser extent, to lactic acid and the intermediates in the high temperatures, long reaction times or the nature of the catalyst may
formation of the latter (glyceraldehyde, dihydroxyacetone and lead to the formation of such products in different amounts [53]. In the
pyruvaldehyde). present study, by qualitative color analysis of the samples (Fig. S6 –
Fructose conversion is known to occur in cascade reactions that are supplementary information), a higher quantity of insoluble products
directed by the presence of Lewis and Brønsted acid sites in the catalyst formed in the presence of Nb2O5 (G or EG), while the smallest amount
structure. The Lewis acid site plays a fundamental role in the isomeri­ was observed when SnO2 (G or EG) was used. When using SnNb (G or
zation of glucose to fructose and on the retro-aldol pathway (forming EG) as the catalyst, the formation of insoluble materials is minimized. In
mainly lactic acid intermediates), and the Brønsted acid sites are more this case, the structural changes in the mixed materials lead to better
active in the dehydration of fructose, generating 5-HMF. Depending on fructose conversion, maintaining the advantage in terms of the forma­
the conversion and desired products, it is necessary to modulate the tion of a minor amount of insoluble products using SnO2 (G or EG).
systems to be employed [24,51,52]. Moreover, the influence of temperature on the fructose conversion
This is clear from examining the trend observed in the catalytic and selectivity for formed products was investigated without a catalyst
evaluation of these systems. In addition to the higher specific surface and in the presence of mixed oxides (SnNb (G) and SnNb (EG)) at 120,
area, mixed oxides SnNb (EG) and SnNb (G) also show a significant 150 and 170 ◦ C after 2 h of reaction (Fig. 7, Fig. S7 and Table T3 -

Fig. 8. Results of reuse for SnNb (G) (A) and SnNb (EG) (B).

68
T.V. dos Santos et al. Catalysis Today 379 (2021) 62–69

supplementary information). As expected, the increase in temperature References


was accompanied by increases in the conversion and the amount of
soluble products formed. In addition, a great increase in insoluble ma­ [1] H.V. Scheller, M. Pauly, D. Loque, Curr. Opin. Plant Biol. 25 (2011) 151–161.
[2] A. Corma, S. Iborra, A. Velty, Chem. Rev. 107 (2007) 2411–2502.
terial formation was observed at 170 ◦ C (see Fig. S8 – supplementary [3] X. Wua, J. Fu, X. Lu, Bioresour. Technol. 119 (2012) 48–54.
information), justifying the use of lower temperatures, such as 150 ◦ C. [4] G. Lv, L. Deng, B. Lu, J. Li, X. Hou, Y. Yang, J. Clean. Prod. 142 (2017) 2244–2251.
Catalyst recyclability is of great importance in biorefinary processes. [5] M. Watanabe, Y. Aizawa, T. Iida, R. Nishimura, H. Inomata, Appl. Catal. A Gen.
295 (2005) 150–156.
In this study, the catalyst was calcined to eliminate organic fractions and [6] S. Souzanchi, L. Nazari, K.T.V. Rao, Z. Yuan, Z. Tan, C.C. Xu, Catal. Today 319
to perform reuse tests for 1 h with 1.5 × 10− 3 g of catalyst at 150 ◦ C. The (2019) 76–83.
results of the recycling experiments are presented in Fig. 8, and no [7] T.V. dos Santos, D.O. da Silva Avelino, M.R. Meneghetti, S.M.P. Meneghetti, Catal.
Commun. 114 (2018) 120–123.
significant variation in the conversion values was observed for SnNb (G) [8] A. Rusanen, R. Lahti, K. Lappalainen, J. Kärkkäinen, T. Hu, H. Romar, U. Lassi,
and SnNb (EG) for each of the 4 cycles. Likewise, previously identified Catal. Today (2019), https://doi.org/10.1016/j.cattod.2019.02.0400. In Press.
products were detected in practically the same amounts in each cycle. [9] J.N. Chheda, Y. Romá-Leshkov, J.A. Dumesic, Green Chem. 9 (2007) 342–350.
[10] H. Kim, Y. Ahn, S. Kwak, Biomass Bioenergy 93 (2016) 243–253.
The liquid samples were collected after the last cycle and analyzed by X-
[11] S. Kassaye, K.K. Pant, S. Jain, Renew. Energy 104 (2017) 177–184.
ray dispersive energy spectroscopy (EDX), and no significant amount of [12] J.B. dos Santos, N.J.A. de Albuquerque, C.L.P.S. Zanta, M.R. Meneghetti, S.M.
tin or niobium was detected (Fig. S9 A). The solid remaining after 4 P. Meneghetti, RSC Adv. 5 (2015) 90952–90959.
reuses was characterized by Raman spectroscopy (Fig. S9 B), and no [13] J.B. dos Santos, F.L. Da Silva, F.M.R.S. Altino, T.S. Moreira, M.R. Meneghetti, S.M.
P. Meneghetti, Catal. Sci. Technol. 3 (2013) 673–678.
significant changes in the catalyst characteristics were observed. [14] X. Wu, J. Fu, X. Lu, Bioresour. Technol. 119 (2012) 48–54.
[15] L. Zhang, G. Xi, K. Yu, H. Yu, X. Wang, Ind. Crops Prod. 98 (2017) 68–75.
4. Conclusions [16] D. Stošića, S. Bennici, V. Rakić, A. Auroux, Catal. Today 192 (2012) 160–168.
[17] J. Guo, S. Zhu, Y. Cen, Z. Qin, J. Wang, W. Fan, Appl. Catal. B 200 (2017) 611–619.
[18] M.M. Antunes, S. Lima, A. Fernandes, J. Candeias, M. Pillinger, S.M. Rocha, M.
The replacement of ethylene glycol by glycerol in pure and mixed F. Ribeiro, A.A. Valente, Catal. Today 195 (2012) 127–135.
metal oxide synthesis using the Pechini method allows the creation of [19] B. Mallesham, P. Sudarsanam, G. Raju, B.M. Reddy, Green Chem. 15 (2013)
478–489.
materials exhibiting higher specific surface areas but did not lead to [20] X. Wang, Y. Song, L. Huang, H. Wang, C. Huang, C. Li, Catal. Sci. Technol. 9 (2019)
significant differences in the amount and nature of the acid sites present 1669–1679.
since all materials exhibit acid sites with weak to moderate strength. [21] M.P. Pechini, US Patent nº 3,330,697, July 11, 1967.
[22] A.E. Danks, S.R. Ball, Z. Schnepp, Mater. Horizons 3 (2016) 91–112.
Moreover, the mixed oxides had more Lewis acid sites, while the number [23] A.L. Quinelato, E. Longo, E.R. Leite, J.A. Varela, Appl. Organomet. Chem. 13
of Bronsted acid sites decreased, leading to an increase in the conversion (1999) 501–507.
when mixed oxide SnNb (G) and SnNb (EG) were used for fructose [24] S.G. Rudisill, S. Shaker, D. Terzic, R. Le Maire, B. Su, A. Stein, Inorg. Chem. 54
(2015) 993–1002.
conversion in aqueous medium in comparison to pure oxides. The
[25] M. Aghbashlo, M. Tabatabaei, H. Rastegari, H.S. Ghaziaskar, J. Clean. Prod. 183
selectivity for 5-HMF and lactic acid (and intermediates) was preserved, (2018) 1265–1275.
regardless of the catalytic system used. The reuse test, which showed [26] M.A. Da Silva, A.S.S. Dos Santos, T.V. Dos Santos, M.R. Meneghetti, S.M.
stability for 4 cycles with no significant reduction in terms of conversion, P. Meneghetti, Catal. Sci. Technol. 7 (2017) 5750–5757.
[27] A. Elci, O. Demirtas, I.M. Ozturk, A. Bek, E. Nalbant Esenturk, J. Mater. Sci. 53
suggested that the hydrophobic characteristics of the rutile structure of (2018) 16345–16356.
SnO2 in contrast to Nb2O5 stability resulted in water-tolerant acid sites. [28] S. Sagadevan, J. Podder, Mat. Res. 19 (2016) 420–425.
[29] R.F. Brandão, R.L. Quirino, V.M. Mello, A.P. Tavares, A.C. Peres, F. Guinhos, J.
C. Rubim, P.A.Z. Suarez, J. Braz. Chem. Soc. 20 (2009) 954–966.
CRediT authorship contribution statement [30] L.J. Burcham, J. Datka, I.E. Wachs, J. Phys. Chem. B 103 (1999) 6015–6024.
[31] F.P. Cardoso, A.E. Nogueira, P.S.O. Patrício, L.C.A. Oliveira, J. Braz. Chem. Soc.
Thatiane V. dos Santos: Conceptualization, Methodology, Investi­ (2012).
[32] D.C. Castro, R.P. Cavalcante, J. Jorge, M.A.U. Martines, L.C.S. Oliveira, G.
gation, Writing - original draft. Dhara B.A. Pryston: Methodology, A. Casagrande, A. Machulek Jr, J. Braz. Chem. Soc. (2015).
Investigation. Geovânia C. Assis: Methodology, Investigation, Writing - [33] R.A. Rani, A.S. Zoolfakar, A.P. O’Mullane, M.W. Austin, K. Kalantar-Zadeh,
original draft. Mario R. Meneghetti: Conceptualization, Supervision. J. Mater. Chem. A. 2 (2014) 15683–15703.
[34] Y. Zhang, J. Wang, J. Ren, X. Liu, X. Li, Y. Xia, G. Lu, Y. Wang, Catal. Sci. Technol.
Simoni M.P. Meneghetti: Conceptualization, Supervision, Writing - 2 (2012) 2485–2491.
review & editing, Project administration. [35] K.S.W. Sing, R.T. Williams, Adsorpt. Sci. Technol. 22 (2004) 773–782.
[36] M. Daturi, L.G. Appel. J. Catal. 432 (2002) 427–432.
[37] A.R. Kamali, D.J. Fray, Adv. Technol. 177 (2012) 819–825.
Declaration of Competing Interest
[38] R.N. Mariammal, K. Ramachandran, B. Renganathan, D. Sastikumar, B Chem. 169
(2012) 199–207.
The authors declare that they have no known competing financial [39] B. Varghese, S.C. Haur, C. Lim, J. Phys. Chem. 112 (2008) 10008–10012.
interests or personal relationships that could have appeared to influence [40] M.L. Savaliya, B.Z. Dholakiya, J. Ind. Eng. Chem. 64 (2018) 352–366.
[41] J. Jehng, I.E. Wachs, Chem. Mater. 3 (1991) 100–107.
the work reported in this paper. [42] A. Yu, R. Frech, J. Power Sources 104 (2002) 97–100.
[43] J. Dijkmans, J. Demol, K. Houthoofd, S. Huang, Y. Pontikes, B. Sels, J. Catal. 330
Acknowledgements (2015) 545–557.
[44] F. de Clippel, M. Dusselier, R. Van Rompaey, P. Vanelderen, J. Dijkmans,
E. Makshina, L. Giebeler, S. Oswald, G.V. Baron, J.F.M. Denayer, P.P. Pescarmona,
This work was supported by the National Council for Scientific and P.A. Jacobs, B.F. Sels, J. M. Chem. Soc. 134 (2012) 10089–10101.
Technological Development (CNPq), the Brazilian Federal Agency for [45] P.Y. Dapsens, C. Mondelli, J. Perez-Ramirez, ChemSusChem 6 (2013) 831–839.
[46] F. Chambon, F. Rataboul, C. Pinel, A. Cabiac, E. Guillon, N. Essayem, Appl. Catal. B
the Improvement of Higher Education (CAPES), the Brazilian Innovation Environ. 105 (2011) 171–181.
Agency (FINEP) and the Alagoas Research Foundation (FAPEAL). TVS, [47] D. Delgado, A. Fernández-arroyo, M.E. Domine, Catal. Sci. Technol. 9 (2019)
DBAP and GCA express their appreciation for fellowships granted by 3126–3136.
[48] M.M. Maronna, E.C. Kruissink, R.F. Parton, F. Soulimani, B.M. Weckhuysen, W.
CAPES and CNPq. SMPM and MRM thank CNPq for research fellowships. F. Hoelderich, Phys. Chem. Chem. Phys. 18 (2016) 22636–22646.
For contributions, the authors thank Prof. Antonio Osimar S. da Silva [49] K. Noda, R.M. De Almeida, N.S. Gonçalves, L.F.D. Probst, O. Sala, Catal. Today 85
(LSCAT-UFAL), Camila B. Dornelas (TecNano-UFAL) and Geraldo Neto (2003) 69–74.
[50] K. Noda, R.M. De Almeida, L.F.D. Probst, N.S. Gonçalves, J. Mol. Catal. A Chem.
(GCAR-UFAL).
225 (2005) 39–46.
[51] C. Sievers, F.L.Y. Noda, L. Qi, E.M. Albuquerque, R.M. Rioux, S.L. Scott, ACS Catal.
Appendix A. Supplementary data 6 (2016) 8226–8307.
[52] M. Moliner, Y. Román-leshkov, M.E. Davis, PNAS 14 (2010) 6164–6168.
[53] G. Tsilomelekis, M.J. Orella, Z. Lin, Z. Cheng, W. Zheng, W. Nikolakis, Green
Supplementary material related to this article can be found, in the Chem. 18 (2016) 1983–1993.
online version, at doi:https://doi.org/10.1016/j.cattod.2020.07.002.

69

You might also like