Photocatalytic conversion of glucose in aqueous suspensions of heteropolyacid–TiO2 composites

M. Bellarditaa, E. I. García-López*a, G. Marcìa, B. Megnab, F. R. Pomillaa and L. Palmisanoa
aSchiavello-Grillone” Photocatalysis Group, Dipartimento di Energia, Ingegneria dell'informazione, e modelli Matematici (DEIM), Università degli Studi di Palermo, Viale delle Scienze Ed. 6, 90128, Palermo, Italy. E-mail: elisaisabel.garcialopez@unipa.it
bDipartimento Ingegneria Civile, Ambientale, Aerospaziale, dei Materiali, Viale delle Scienze, 90128 Palermo, Italy

Received 26th May 2015 , Accepted 1st July 2015

First published on 1st July 2015


Abstract

Commercial and home prepared TiO2 samples were functionalized with a commercial Keggin heteropolyacid (HPA) H3PW12O40 (PW12) or with a hydrothermally home prepared K7PW11O39 salt (PW11). All the materials were characterized by specific surface area measurements (BET), XRD analyses, Raman, DRS along with SEM observations and they have been used for glucose photocatalytic conversion in an aqueous suspension. Different reaction extents and distribution of intermediate oxidation products were observed depending on the photocatalyst. Gluconic acid, arabinose, erythrose and formic acid were observed as oxidation products when bare TiO2 or HPA/TiO2 composite materials were used. Glucose isomerization to form fructose was also observed and in some runs traces of glucaric acid and glyceraldehyde were also found. The carbon mass balance was accomplished in the presence of the commercial Evonik P25 TiO2 powder and the composites where TiO2 was present, whereas the presence of the solvothermically prepared TiO2 gave rise to a carbon unbalance, due to strong adsorption of the products on the photocatalyst surface. No reactivity was observed in the presence of PW12 alone while PW11 induced only isomerization of the glucose.


1. Introduction

The search for alternative resources for the synthesis of chemicals currently produced from non-renewable sources has directed the activities of researchers towards the use of different raw materials such as biomass.1 It seems particularly interesting to use lignocellulose (cellulose, hemi-cellulose and lignin) which can derive from agricultural wastes. Glucose, obtained from cellulose, can be used for the sustainable production of high value chemicals. To this aim, catalytic processes at high pressure and temperature, pyrolysis, gasification or conversion under supercritical conditions have been the object of scientific research. Glucose can be used to obtain ethanol by fermentation,2 sorbitol and mannitol by hydrogenation,3 5-hydroxymethyl furfural by dehydrocyclization4 and also to produce hydrogen.5 Glucose is the monosaccharide most extensively studied in oxidation reactions particularly to obtain gluconic and glucaric acids.6 This last reaction and in general, the selective oxidation of alcohols to their corresponding carbonyl compounds has attracted attention in the field of catalysis, due to its strategic importance.7 Gluconic acid, with an annual estimated market of 6 × 104 ton, is used as a biodegradable chelating agent, a water soluble cleansing agent and an intermediate in food and pharmaceutical industries.8 It is currently industrially prepared by fermentation of glucose by Aspergillus niger,9 although this process presents some drawbacks, as the disposal of dead microbes and the slow reaction rate. The heterogeneous catalytic oxidation of glucose has been presented as an attempt to overcome the problems of the biological process. The heterogeneous catalytic oxidations of sugars are performed by using supported noble metal catalyst in aqueous medium with batch reactors in the presence of air or oxygen under atmospheric pressure at temperatures of 293–353 K. The reaction is carried out at almost neutral or basic pH's (7–9) in order to allow the carboxylate anions to desorb from the catalyst surface avoiding their degradation.10 The mild reaction conditions, along with the fact that the reactants are renewable and the products are environmentally benign because of their biodegradability, make the catalytic oxidation of carbohydrates a paradigm of green chemistry. Metal catalysts as Pt, Pd, Rh, Bi or Pb supported on TiO2, Al2O3 or activated carbon have been used as catalysts but Au seems to be the most promising one.11–15 Gold nanoparticles supported on activated carbon16 or ZrO2 (ref. 17) showed good catalytic activity for oxidation of glucose to gluconic acid at 50 °C in the presence of O2 at pH's ranging between 8 and 10.

In this context, heterogeneous photocatalysis can be also considered as an alternative. The heterogeneous photocatalytic technology by using semiconductor oxides as photomediators is known as a process suitable to degrade organic and inorganic pollutants both in vapour and in liquid phases under very mild experimental conditions.18 It is generally accepted that TiO2 is the most reliable photocatalyst19 and Colmenares et al. investigated the glucose photocatalytic oxidation in the presence of TiO2 in acetonitrile–water suspensions.20–22 They obtained glucaric and gluconic acids along with arabitol and reported that the presence of acetonitrile stabilized the carboxylic acids by solvation suppressing their further oxidation.22 It is rare to obtain acceptable selectivity values for partial photocatalytic oxidation reactions in the presence of only water as the solvent.23 Chong et al. studied the conversion of glucose under anaerobic conditions in TiO2-rutile aqueous suspensions and they found arabinose, erythrose and hydrogen as the products.24 Heteropolyacid (HPA) clusters have been studied as homogeneous photocatalysts, due to their ability to absorb UV light. The absorption of light by the ground electronic state of the HPA produces a charge transfer-excited state HPA* which can behave as a better oxidant species than the corresponding ground states.25 Under irradiation with light of suitable wavelengths, HPA reduces to HPA, the so called “heteropoly-blue”. The heteropoly-blue species is relatively stable, absorbs visible light and is readily reoxidized to the original HPA. This process can occur both with the plenary HPA Keggin species (H3PW12O40) and with the lacunary Keggin cluster (K7PW11O39).26 HPA photo-reduction has been proved to be synergistically enhanced in HPA/TiO2 composites where photo-generated electrons can be transferred from the conduction band of TiO2 to HPA. In this way the charge-pair recombination in the TiO2 is delayed.27,28 Heteropolyacids, such as H3PW12O40, are also strong acid catalysts able to catalyze at low temperatures a wide range of catalytic processes.29 They exhibit very strong Brönsted type acidity, making them suitable for various acidic reactions, such as esterification, transesterification, hydrolysis, Friedel–Crafts alkylation and acylation and Beckmann rearrangement.30

The present paper reports the preparation of nanometer-sized TiO2 particles by a solvothermal method. The commercial saturated H3PW12O40 (labelled as PW12) and the home prepared lacunary monovacant K7PW11O39 Keggin salt (labelled as PW11) were coupled by a solvothermal treatment with TiO2 obtaining PW12/TiO2 and PW11/TiO2 composite materials. Moreover, also the impregnation of the saturated Keggin unit PW12 on commercial TiO2 surface was performed for the sake of comparison. Some physico-chemical properties of the prepared materials were investigated along with their photoactivity for glucose oxidation in aqueous medium at natural pH. The experiments were carried out under mild conditions: room temperature, atmospheric pressure in aqueous suspensions and by using an inexpensive material. Photocatalysis by using heterogenized heteropolyacids is a novel field, and the glucose partial oxidation by using these solids has never been investigated before, to the best of our knowledge.

2. Experimental

2.1. Photocatalysts preparation and characterization

A first set of powders was obtained by wet impregnation of commercial TiO2 (Evonik P25) with a solution of a commercial heteropolyacid (HPA), i.e. tungstophosphoric acid H3PW12O40 (Aldrich reagent grade 99.7%), labelled as PW12. In particular TiO2 (8.3 g) was added to a water solution (50 mL) containing the appropriate amount of PW12 (2.3 g). The suspension was stirred for ca. 1 h and then it was divided in two parts. One of them was hydrothermally treated in a teflonated autoclave for 48 hours at 200 °C and the obtained white powder filtered and dried at 60 °C. This sample has been named PW12/P25 solv. The other part of the suspension was, instead, evaporated until dryness in a vacuum-dryer apparatus and the obtained powder labelled as PW12/P25.

An alternative method was followed to obtain TiO2 by using titanium isopropoxide, Ti(OPr)4, (Aldrich 97%) as the precursor. The composite materials were prepared by adding the alkoxide precursor (32 mL) to the PW12 (2.3 g) aqueous solution (50 mL) and the resulting suspension was subjected to a hydrothermal treatment at 200 °C for 48 h (in this case the system reached a pressure of ca. 17 atm). The resulting bluish powder was washed three times with hot water and finally filtered and dried at 60 °C. This sample was denoted as PW12/TiO2 solv. The analogous bare TiO2 was prepared under the same experimental conditions in the absence of PW12 and labeled as TiO2 solv.

Another set of samples was prepared by using a monolacunary PW11 Keggin salt. The heteropolyacid K7PW11O39 has been obtained by following the Haraguchi method.31 20 g of commercial H3PW12O40·26H2O were dissolved in 100 mL of hot water, then 1.0 g of KCl was added and the pH of the solution adjusted to 5 with KHCO3 1 M. The obtained solid was filtered and dried at room temperature. For the preparation of the K7PW11O39/TiO2 materials, 3.6 mL of titanium isopropoxide was dissolved in 24 mL of 2-propanol under stirring at room temperature for 1 h. 0.125 or 0.250 g of K7PW11O39 were dissolved in 2 mL of hot water and then added, under vigorous stirring, to the alcoholic solution of the TiO2 precursor. The resulting suspension was adjusted to pH 5 with acetic acid 1 M, transferred to the teflonated autoclave and heated at 433 K for 48 h. The white bluish powder obtained was washed with water and eventually dried at room temperature. The obtained powders were named PW11-X/TiO2 solv (where X = 0.125 or 0.250 g, depending on the amount of PW11 used).

Bulk and surface characterizations were carried out in order to define some physicochemical properties of the powders. Their crystalline phase structure was determined at room temperature by powder X-ray diffraction analysis (PXRD) carried out by using a Panalytical Empyrean, equipped with CuKα radiation and PixCel1D (tm) detector. The specific surface areas (SSA) were determined in a Flow Sorb 2300 apparatus (Micromeritics) by using the single-point BET method. Scanning electron microscopy (SEM) was performed using a FEI Quanta 200 ESEM microscope, operating at 20 kV on specimens upon which a thin layer of gold had been evaporated. An electron microprobe used in an energy dispersive mode (EDAX) was employed to obtain information on the actual metals content present in the samples. Raman spectra were obtained by means of a BWTek-i-micro Raman Plus System, equipped with a 785 nm diode laser. The measurements were performed focusing the sample by a 20× magnification lens, spot size was around 50 μm. The accuracy of Raman shift was around 3 cm−1. The power of the laser used was 15% of the maximum value that was around 300 mW. Infrared spectra of the samples in KBr (Aldrich) pellets were obtained with a FTIR-8400 Shimadzu spectrophotometer and the spectra were recorded with 4 cm−1 resolution and 256 scans. The diffuse reflectance spectra (DRS) were recorded in air at room temperature in the 250–800 nm wavelength range using a Shimadzu UV-2401 PC spectrophotometer, with BaSO4 as the reference material.

2.2. Photocatalytic activity

The photoreactivity runs were carried out at room temperature and ambient pressure in a 800 mL open reactor irradiated in the UV region with an immersed 125 W medium pressure Hg lamp (Helios Italquartz, Italy). The initial aqueous glucose concentration was 1 mM and the runs were carried out at natural pH. The impinging radiation intensity was measured by a radiometer Delta Ohm DO9721, and it was 5.5 mW cm−2. The amounts of photocatalyst needed to absorb all the photons emitted by the lamp in the reacting suspension was checked by using the same radiometer and depended on the catalyst used. In particular, in some cases 0.3 g L−1 of catalyst were sufficient (see TiO2 P25, PW12/P25 solv and PW12/P25), whereas in other cases the amount was 2.0 g L−1 (see TiO2 solv, PW11-0.125/TiO2 solv and PW11-0.250/TiO2 solv) or 2.8 g L−1 (see PW12/TiO2 solv).

Air was not bubbled during the experiments but the vessel was just opened in ambient conditions. The values of substrate concentration before the addition of catalyst and before the starting of irradiation were measured in order to determine the substrate adsorption on the catalyst surface under dark conditions. During the photoreactivity runs samples were withdrawn at fixed times and immediately filtered through 0.2 μm membranes (HA, Millipore) before analyses. The quantitative determination and identification of glucose and its degradation products were performed by means of a Thermo Scientific Dionex ultimate 3000 HPLC equipped with a Diode Array and refractive index detectors. The column was a REZEK ROA Organic acid H+ phenomenex, the eluent an aqueous 2.5 mM H2SO4 solution and the flow rate 0.6 mL min−1. Retention times and UV spectra of the compounds were compared with those of standards purchased from Sigma-Aldrich with a purity of >99%. All of the runs lasted ca. 6 h and were performed at least twice.

In order to test the adsorption extent of some reaction intermediates (arabinose and gluconic acid) on the different photocatalysts used, some adsorption tests were carried out following the procedure below reported. An amount of 2.8 g L−1 of photocatalyst was dispersed in an aqueous solution containing 1 mmol L−1 of arabinose or 1 mmol L−1 of gluconic acid. The suspension was maintained under stirring in dark condition for 6 h. 5 ml of suspension were withdrawn every hour and the concentrations of arabinose or gluconic acid were analysed after the separation of the photocatalyst.

3. Results and discussion

3.1. Characterization of the photocatalysts

Table 1 reports some physicochemical features of the powders investigated as photocatalysts. The specific surface area (SSA) of TiO2 sample prepared under solvothermal conditions was 260 m2 g−1, a value much higher than that of the commercial TiO2 Evonik P25 (50 m2 g−1). The SSA's of the commercial PW12 and the home prepared PW11 were 15 and 80 m2 g−1, respectively, and as a general trend, those of all of the composite materials were smaller than those of the bare TiO2 samples.
Table 1 Some physicochemical data concerning the characterization of the TiO2 and HPA/TiO2 photocatalysts
Photocatalyst S.S.A. [m2 g−1] Egap [eV] Nominal EDAX
W atomic [%] Ti atomic [%] W atomic [%] Ti atomic [%]
a Some agglomerates in the PW12/P25 sample present values of W atomic percentage and Ti atomic percentage of ca. 30% and ca. 70%, respectively.
TiO2 P25 50 3.2
PW12 15 3.2
PW12/P25 48 3.0 9 91 10 ± 1a 90 ± 1a
PW12/P25 solv 43 3.0 7 93 7 ± 1 93 ± 1
TiO2 solv 260 3.2
PW12/TiO2 solv 176 3.2 7 93 10 ± 0.1 90 ± 0.1
PW11 80 3.4
PW11-0.125/TiO2 solv 206 3.1 4 96 5 ± 1 9 ± 1
PW11-0.250/TiO2 solv 196 3.2 7 93 7 ± 1 93 ± 1


XRD diffractograms of all of the prepared samples are reported in Fig. 1. In Fig. 1(A) the diffractograms correspond to the bare PW12, commercial TiO2 Evonik P25 (in the following named only TiO2 P25) and the composites obtained with these two substances. The PW12 presents a crystalline structure and the commercial TiO2 P25 consists of anatase and rutile polymorphs. In the composite materials, new peaks, attributable to the heteropolyacid, in addition to those of TiO2 are present and they are less intense in the PW12/P25 solv sample than in the PW12/P25 one. This finding can account for the best dispersion of HPA in the material prepared solvothermically. Indeed, the more defined peaks can be due to the heterogeneity of the dispersion of PW12 in the P25 impregnated sample, as confirmed by SEM (see in the following). Fig. 1(B) shows the diffractograms of PW12 along with the home prepared bare TiO2 and the composite PW12/TiO2 solv prepared solvothermically. In the diffractogram it can be noticed the presence of anatase phase for both bare TiO2 and PW12/TiO2 solv samples without significant differences and without the presence of peaks ascribable to HPA. This result suggests a good statistic mixing of HPA with TiO2. Fig. 1(C) reports the diffractograms of PW11, TiO2 solv and PW11-X/TiO2 solv samples (see Experimental section for the meaning of the code). The diffractogram of bare PW11 indicates a good crystallinity of this sample although the characteristic peaks have not been reported in the literature. Both PW11-X/TiO2 solv samples do not present a strong evidence of the PW11 crystalline phase segregated. The only significant difference between PW11-0.125/TiO2 solv and PW11-0.250/TiO2 solv samples and TiO2 solv is the wide peak localized at 2θ = 14°.


image file: c5ra09894g-f1.tif
Fig. 1 XRD patterns of the photocatalysts: (A) (a) PW12; (b) TiO2 P25; (c) PW12/P25 solv; (d) PW12/P25; (B) (a) TiO2 solv; (b) PW12/TiO2 solv and (C) (a) PW11; (b) TiO2 solv; (c) PW11-0.125/TiO2 solv; (d) PW11-0.250/TiO2 solv (*Rutile; + Anatase).

The samples were also investigated by FTIR (spectra not reported for the sake of brevity) to check the structural integrity of the Keggin unit after the preparation of the HPA/TiO2 composites. The arrangement structure of the PW12 consists of a PO4 tetrahedron surrounded by four W3O9 groups formed by edge sharing octhaedra.32 This arrangement gives rise to four stretching bands: P–O stretching mode at 1080 cm−1, W[double bond, length as m-dash]O stretching observed at 990 cm−1 and two peaks at ca. 910 cm−1 and 810 cm−1 attributed to two types of W–O–W units.33 It is difficult to characterize the Keggin structure in the composites by IR spectroscopy because some of the bands overlap with those assigned to TiO2 powder; the latter, in fact, presents an intense and broad vibration band originated from Ti–O–Ti bonds located at wavenumbers lower than 900 cm−1.

Raman spectroscopy allowed a much better understanding of the structural modifications induced at the surface. In Fig. 2(A) both bare TiO2 samples (TiO2 solv and TiO2 P25) show Raman peaks centered at 144, 197, 399, 513, and 639 cm−1 attributable to the Eg, Eg, B1g, A1g and B2g modes of anatase TiO2.34 On the contrary, the characteristic peaks of the rutile phase, that should be located at 444 and at 609 cm−1, are not observed in the samples, as reported before,27 probably because the rutile crystallites observed by XRD are located in the bulk of the material.


image file: c5ra09894g-f2.tif
Fig. 2 Raman spectra of the samples: (A) bare TiO2 and PW12 containing materials, and (B) PW11 containing samples.

The Raman spectrum of PW12 shows a sharp and intense band at 1008 cm−1 and a peak at 990 cm−1 assigned to P–O vibrations and bands at lower wavenumbers, attributed to W–O (925 cm−1) and W–O–W (880 cm−1) vibrations.33

In the HPA/TiO2 composites the Raman peaks attributable to anatase phase are also present indicating that the crystalline form is preserved on the surface after the introduction of the heteropolyacid. The four HPA characteristic bands are not present due to the very low amount of HPA in the samples. However, the sharper and more intense band at 1008 cm−1 and the shoulder at 990 cm−1 are present for the PW12/P25 composite in Fig. 2(A) and no significant shift can be observed. On the contrary, for the samples treated by the solvothermal method, PW12/P25 solv and PW12/TiO2 solv, the 1008 and 990 cm−1 peaks appear as a unique broad peak shifted to ca. 960 cm−1. These finding can be attributed to an interaction between the oxygen atom of the Keggin anion and the hydroxyl groups on the TiO2 surface.35 As far as the lacunary Keggin is concerned, Fig. 2(B) shows the high complexity of the Raman spectrum of this HPA. The spectra of the binary PW11/TiO2 samples confirm the presence of anatase along with a small shoulder at ca. 940 cm−1 due to the presence of PW11 in the composites (see the intense band observed for the bare PW11 at 960 cm−1).

SEM microphotographs of some selected materials are reported in Fig. 3(A) and (B). In particular, Fig. 3(A) reports some pictures of the bare PW11 (a–d) and of the home prepared TiO2 (e–f) samples whereas Fig. 3(B) reports some pictures of composite powders. The morphology of the PW11 salt appears completely different compared with that of the other samples. This sample seems consisting of very large crystals surrounded by others growing small crystals. On the contrary, the TiO2 sample which was solvothermally prepared consists of agglomerates of primary particles (ca. 40–60 nm), whose size ranges between 2.5 and 30 μm. The PW11-0.125/TiO2 solv, PW11-0.250/TiO2 solv and PW12/TiO2 solv composite samples (Fig. 3(B)a–f, respectively) appear very similar to the bare TiO2 sample (Fig. 3(A)e and f), indicating that the small content of PW11 or PW12 did not modify the morphology of the majority component TiO2. From the perusal of Fig. 3(A) and (B) it can be concluded that in the case of the PW11-0.250/TiO2 solv the size of the primary particles resulted smaller (ca. 20 nm). Consequently, it seems that the presence of a higher amount of PW11 caused a decrease of the size of the primary particles. The morphology of PW12/P25 (Fig. 3(B)g and h) and PW12/P25 solv (not reported in Fig. 3), is very similar to the bare material (SEM picture of bare TiO2 P25 has been already reported28). The agglomerates of these particles present the same shape and consist of nanoparticles with similar sizes (ca. 50 nm). Table 1 reports the nominal and the average EDAX values of atomic percentage of W in PW11 or PW12 and the atomic percentage of Ti in the TiO2. EDAX measurements confirmed a homogeneous content of the HPAs onto the catalyst with the exception of the PW12/P25 sample where tungsten was present in some agglomerates in much higher content with respect to the nominal one. In all of the other samples, the measured amount of W and Ti was always very close to the nominal one. Fig. 4 (A) and (B) report the diffuse reflectance UV-Vis spectra (DRS) of the bare TiO2 and HPA/TiO2 composite samples. The insets report the absorbance spectra obtained by applying the Kubelka–Munk function, F(R), to the diffuse reflectance spectra. All of the spectra are characterized by a charge transfer process, from O 2p to Ti 3d for TiO2 or to HOMO–LUMO transition for the HPAs.


image file: c5ra09894g-f3.tif
Fig. 3 SEM microphotographs of (A) two agglomerates (a–d) of bare PW11 and (e and f) bare TiO2 at different magnifications and (B) (a–h) some selected composite materials at two different magnifications: (a and b) PW11-0.125/TiO2 solv; (c and d) PW11-0.250/TiO2 solv; (e and f) PW12/TiO2 solv; (g and h) PW12/TiO2 P25.

image file: c5ra09894g-f4.tif
Fig. 4 Diffuse reflectance spectra of the samples: (A) (a) TiO2 P25; (b) TiO2 solv; (c) PW12/P25; (d) PW12/P25 solv; (e) PW12/TiO2 solv; (f) PW12 and (B) (a) PW11; (b) PW11-0.125/TiO2 solv, and (c) PW11-0.250/TiO2 solv. The insets report the absorbance spectra obtained by applying the Kubelka–Munk function, F(R), to the diffuse reflectance spectra.

Fig. 4(A) reports the spectra of the bare TiO2 and PW12/TiO2 solv samples, whereas in Fig. 4(B) those concerning the samples prepared with the lacunary HPA are shown. The spectrum of PW11 evidences a higher energy for the transition HOMO–LUMO compared to that of PW12. The Kubelka–Munk function F(R) has been used to obtain the Tauc plots36 where the extrapolation in the linear fitting of the plot (F(R)E)1/2 vs. incident light energy in eV gives the band gap energy (see Table 1). The presence of the HPA gave rise to a slight increase of the band gap energy, particularly where the PW11 was present. Based on the above physico-chemical characterization results, it can be concluded that the primary Keggin structure of the saturated and lacunary HPA remained virtually unchanged after the deposition of the cluster on the oxide surface. Different kind of interactions between the saturated or lacunary Keggin unit and the TiO2 surface can be hypothesized. For the PW12/TiO2 solv composite, it can be suggested that the saturated Keggin unit interacts with TiO2 by hydrogen bonding and acid–base reaction as reported before.35 On the contrary, according to Ma et al., the removal of a tungsten–oxygen octahedral from a saturated PW12O403− gives rise to the lacunary anion (PW11O397−) that results highly nucleophilic and can react easily with electrophilic groups such as titanium atoms of Ti-OH groups present in TiO2. Therefore, in the PW11/TiO2 composite, K7PW11O39 presents vacant sites, which allow connecting two TiO4 units of the TiO2 network to make up tungsten–oxygen octahedral lacunas. Consequently, the terminal nucleophilic oxygen atoms of the K7PW11O39 become bridge atoms that allow to connect PW11 and TiO2 via W–O–Ti bonds.37

3.2. Glucose photocatalytic conversion

Fig. 5–7 report the results of the photocatalytic glucose conversion in the presence of different catalysts. Fig. 5 reports the activity of the bare TiO2 powders. By using TiO2 P25, the decrease of glucose concentration was accompanied by the appearance of various compounds, mainly formic acid and arabinose. Other oxidation products as erythrose and gluconic acid were also formed in lower amounts. Fructose, a glucose isomerization product, has been also found. It is worth to mention that the amount of P25 sufficient to absorb all the photons emitted by the lamp was 0.3 g L−1, under the experimental conditions used. TiO2 solv possesses a lower ability in absorbing photons (this feature can be related to the higher granulometry of the powder) and 2 g L−1 were necessary to absorb all the emitted photons. Consequently, two runs with TiO2 solv were carried out: the first one with 2 g L−1, the second one with 0.3 g L−1. As reported in Fig. 5, for the run with 2 g L−1 of photocatalyst, the decrease of glucose concentration was faster than in the presence of TiO2 P25, but the amount of formic acid was much lower. Arabinose was found as a product, along with low amounts of erythrose and gluconic acid. In this case, the isomerization product fructose was also observed. By employing a lower amount of TiO2 powder (0.3 g L−1) the glucose conversion decreased, the quantities of formic acid, erythrose and gluconic acid slightly decreased and arabinose was completely absent. Only the fructose concentration increased moderately.
image file: c5ra09894g-f5.tif
Fig. 5 Evolution of glucose and photocatalytic reaction products versus irradiation time in the presence of bare TiO2 samples. (●) Glucose, (▲) fructose, (▼) erythrose, ([pentagon filled]) arabinose, (◆) gluconic acid, and (■) formic acid. The average oscillation percentage of the experimental data was ca. ± 2%.

image file: c5ra09894g-f6.tif
Fig. 6 Evolution of glucose and photocatalytic reaction products versus irradiation time in the presence of the binary materials composed of PW12 and TiO2. (●) Glucose, (▲) fructose, (▼) erythrose, ([pentagon filled]) arabinose, (◆) gluconic acid, ([hexagon filled, flat-side down]) glucaric acid, and (■) formic acid. The average oscillation percentage of the experimental data was ca. ± 2%.

image file: c5ra09894g-f7.tif
Fig. 7 Evolution of glucose and photocatalytic reaction products versus irradiation time in the presence of the binary materials composed of PW11 and TiO2. (●) Glucose, (▲) fructose, (▼) erythrose, ([pentagon filled]) arabinose, (◆) gluconic acid, ([hexagon filled, flat-side down]) glucaric acid, and (■) formic acid. The average oscillation percentage of the experimental data was ca. ± 2%.

Table 2 reports conversion and selectivity to the different species found along with the carbon mass balance calculated after 6 hours of irradiation for the experiments reported in Fig. 5. The glucose conversion, X, and the selectivity, S, to fructose, gluconic acid, arabinose and erythrose have been calculated as follows:

 
X = ([glucose]i − [glucose])/[glucose]i × 100 (1)
 
S = [product]/([glucose]i − [glucose]) × 100 (2)
where [glucose]i is the initial glucose concentration, [glucose] the concentration after 6 hours of irradiation and [product] denotes the product concentration analysed at the same time. It is important to highlight that the carbon atom mass balance has been satisfied when bare TiO2 P25 has been used but not in the presence of the bare home prepared TiO2 solv sample. In fact, the color of the latter sample turned pale orange after all of the runs, indicating that some species remained strongly adsorbed on its surface. The greater ability of TiO2 solv sample to adsorb the reaction intermediates compared to that of TiO2 P25 sample has been confirmed by adsorption tests carried out in the presence of arabinose and in the presence of gluconic acid chosen as representative intermediates. These tests indicate that the TiO2 solv sample was able to adsorb the compounds above mentioned ca. seven times more than TiO2 P25 sample.

Fig. 6 reports the evolution of glucose and reaction products in photocatalytic experiments carried out in the presence of PW12/TiO2 composites. The samples obtained by impregnation of TiO2 P25 with PW12, i.e. PW12/P25 and PW12/P25 solv, convert glucose faster than the correspondent bare TiO2 P25. In both cases fructose was the main species observed during the run, but also a significant amount of gluconic acid was formed, particularly in the presence of PW12/P25 solv. It is worth noting that the amount of gluconic acid formed by using these two composite materials was always higher than that observed in the presence of bare TiO2 P25. Small amounts of formic acid and erythrose were also found along with traces of glucaric acid. By using PW12/TiO2 solv as the photocatalyst (2.8 g L−1 were needed to absorb all the photons emitted by the lamp), the conversion of glucose was lower with respect to PW12/P25 and PW12/P25 solv. Fructose resulted to be the main species, along with gluconic acid and formic acid. Table 3 collets the X, S and B data obtained for the runs reported in Fig. 6. The presence of PW12 along with TiO2 P25, in both samples obtained by impregnation or by solvothermal treatment, caused an increase in the glucose conversion, the absence of arabinose and a strong decrease of formic acid and erythrose in comparison to TiO2 P25 (see Table 2).

Table 2 Photoreactivity assessment of bare TiO2 powders: glucose conversion, X, selectivity to the identified reaction products, S, and carbon mass balance, B, after 360 minutes of irradiation
    TiO2 P25 (0.3 g L−1) TiO2 solv (0.3 g L−1) TiO2 solv (2 g L−1)
X [%] Glucose 41 ± 1 40 ± 1 72 ± 1
S [%] Fructose 14 ± 1 20 ± 1 6 ± 0.5
Gluconic acid 0.6 ± 0.1 3 ± 0.2 1 ± 0.1
Arabinose 42 ± 2 26 ± 1
Erythrose 28 ± 1.5 13 ± 1 5 ± 0.5
B [%] Carbon 99 ± 1 75 ± 2 55 ± 1


Table 3 Photoreactivity assessment of composites containing PW12 and TiO2: glucose conversion, X, selectivity to the identified reaction products, S, and carbon mass balance, B, after 360 minutes of irradiation
    PW12/P25 (0.3 g L−1) PW12/P25 solv (0.3 g L−1) PW12/TiO2 solv (2.8 g L−1)
X [%] Glucose 64 ± 1 85 ± 1 39 ± 1
S [%] Fructose 68 ± 2 55 ± 2 56 ± 2
Gluconic acid 22 ± 1 34 ± 1 20 ± 1
Glucaric acid 1.5 ± 0.2
Erythrose 11 ± 1
B [%] Carbon 97 ± 1 99 ± 1 92 ± 1


On the contrary, conversion to fructose and gluconic acid increased. In particular, in the presence of PW12/P25 solv the amount of valuable gluconic acid was the highest obtained in this work. Notably this last sample did not give rise to leaching of PW12 in the liquid phase, contrary to what observed for PW12/P25 prepared by impregnation.

In the light of the results above presented, some general considerations can be done. The presence of PW12 in PW12/P25 and PW12/P25 solv samples, resulted beneficial. On the contrary, PW12/TiO2 solv composite showed a lower reactivity than that obtained in the presence of PW12/P25 samples giving rise mainly to fructose. This finding can be explained by taking into account the different preparation methods used to obtain this composite in comparison to those containing P25.

In the latter cases, PW12 was added to TiO2 already crystallized, whereas PW12/TiO2 solv sample was prepared by adding the Ti organometallic precursor to the PW12 solution; consequently, PW12 was dispersed not only on the TiO2 surface but mainly in the bulk. This fact has been evidenced by XRD investigations where the pattern of the PW12/TiO2 solv does not indicate the presence of the heteropolyacid. Consequently, it is evident that the presence of PW12 on the TiO2 P25 surface, both prepared by impregnation or solvothermically, was able to modify the reactivity and selectivity of the bare TiO2 towards the formation of less degraded products in the glucose oxidation process. Also in the PW12/TiO2 solv sample the presence of PW12 in the bulk of TiO2 influenced the reactivity of TiO2 giving rise to a decrease in the formation of formic acid and an increase of fructose.

Moreover, the presence of PW12 favoured the products desorption from the photocatalyst surface making possible the achievements of the carbon mass balance contrarily to what observed with the bare TiO2 solv sample. Adsorption tests carried out as those discussed for TiO2 P25 and TiO2 solv sample indicated that the presence of PW12 drastically reduced the adsorption ability of TiO2 solv. This finding was also observed in the presence of PW11, as it will be discussed later. It is worth to mention that both HPA species, PW11 and PW12, are soluble in water. PW12 used as photocatalyst in homogeneous regime did not show to be photoactive; conversely, PW11 gave rise to a slight conversion of glucose to fructose. The products formed in the presence of samples consisting of the PW11 lacunary cluster and TiO2 (PW11-0.125/TiO2 solv and PW11-0.250/TiO2 solv) are reported in Fig. 7. Also for these two samples different amounts were used (0.3 g L−1 and 2 g L−1), due to their scarce ability to absorb all the photons emitted by the lamp (Table 4).

Table 4 Photoreactivity assessment of composites containing PW11 and TiO2: glucose conversion, X, selectivity to the identified products, S, and carbon mass balance, B, after 360 minutes of irradiation
    PW11 (0.3 g L−1) PW11-0.125/TiO2 solv (2 g L−1) PW11-0.250/TiO2 solv (2 g L−1) PW11-0.125/TiO2 solv (0.3 g L−1) PW11-0.250/TiO2 solv (0.3 g L−1)
X [%] Glucose 20 ± 1 44 ± 2 53 ± 2 20 ± 1 28 ± 1
S [%] Fructose 99 ± 1 24 ± 1 31 ± 1 34 ± 1 24 ± 1
Gluconic acid 4 ± 0.2 19 ± 1 9 ± 0.5 35 ± 2
Glucaric acid 1 ± 0.1
Arabinose 38 ± 2
Erythrose 37 ± 1 19 ± 1 23 ± 1 16 ± 1
B [%] Carbon 99 ± 1 99 ± 1 90 ± 1 95 ± 1 95 ± 1


No leaching of PW11 occurred during the experiments, indicating that PW11 played a role exclusively in heterogeneous phase. In the runs carried out with 0.3 g L−1, the materials were scarcely active (see Fig. 7) and some decrease in the concentration of glucose was observed along with the formation of small amounts of fructose, gluconic acid, erythrose and formic acid. The use of 2 g L−1 of photocatalyst gave rise to a high increase of reactivity. PW11-0.250/TiO2 solv converted glucose faster than PW11-0.125/TiO2 solv but in both cases the most important product was formic acid that was surprisingly found in higher amount in the presence of the latter sample. In both systems, the presence of gluconic acid, erythrose and fructose was also detected. The reactivity of this two photocatalysts differed only because arabinose was found in the presence of PW11-0.125/TiO2 solv while small amounts of glucaric acid were detected in the PW11-0.250/TiO2 solv. For the composite PW11/TiO2 solv samples the conversion of glucose was lower than for TiO2 solv, whereas the selectivity to gluconic acid higher. The largest amount of photocatalyst (2 g L−1 vs. 0.3 g L−1) caused a higher conversion of glucose but also a higher formation of formic acid.

It is worth noting that the presence of PW11 in the composite PW11/TiO2 solv samples favoured the desorption of the products from the photocatalyst surface, however it caused a strong increase of the overoxidation product, i.e. gluconic acid and formic acid.

On the basis of the previous considerations we can conclude that the presence of HPA (PW12 or PW11) changed the glucose reaction mechanism; indeed, in the runs carried out by using bare TiO2 the main products were arabinose and erythrose, whereas fructose and gluconic acid were preferentially formed when the PW12 composites were used as photocatalysts. When the PW11 composites were employed, the formation of formic acid was favoured. Probably, HPA modified the surface properties of TiO2 and in particular the type and the strength of acid site along with their distribution. This finding can justify the higher amounts of fructose obtained when composite catalysts containing PW12 were used. Indeed, in the literature it is reported that the isomerization of glucose to fructose is catalysed by Lewis acids.38 In particular, this effect was evident in the case of P25/TiO2 samples where HPA is preferentially localized on the TiO2 surface. The presence of PW12 was able to change the reaction mechanism, probably because it modified the acid properties of the catalyst surface. This finding indicate that PW12 was able to explicate its acid function (see the higher formation of fructose) reducing the oxidizing ability of TiO2 (see the disappearance of arabinose and erythrose and the formation of gluconic acid). Indeed, when TiO2 was used alone, according to Chong et al.24 the most favoured reactions were due to the α scission (C1–C2 cleavage) giving rise to the formation of arabinose, erythrose and glyceraldehyde (this last observed only in traces). On the other hand, the oxidation of glucose to gluconic acid (as first step) and to glucaric acid (as second step) were observed only during the runs carried out by using PW12 containing composite photocatalysts. On the contrary, the presence of PW11, that does not show any acid function, induces an increase of the oxidizing ability of TiO2, probably acting through the formation of the heteropoly-blue species26 and consequently reducing the electron–hole recombination on TiO2. However, it is worth noting that the sample PW11-0.125/TiO2 solv is more oxidizing than the sample PW11-0.250/TiO2 solv. This fact can be explained by considering that an increase of the amount of PW11 over a certain value could cover the TiO2 surface or favour the electron–hole recombination (two effects that in any case are detrimental for the reactivity). It is worth to remark that an important feature in the catalytic oxidation of carbohydrate molecules is the regio- and chemoselectivity, but these aspects have been considered out of the scope of this work. In Scheme 1 it has been summarized a possible reaction sequence by considering the Hoffman structures of the analysed molecules in the D-form.


image file: c5ra09894g-s1.tif
Scheme 1 Hypothesis of reaction sequence for the photocatalytic glucose oxidation.

The oxidation of glucose can occur by the oxidation of the anomeric center (C1) giving rise to gluconic acid. Successively, an oxidative attack to the C2 carbon gives rise to a formation of formic acid and arabinose. A further oxidative attack to the C2 carbon in arabinose molecule releases another formic acid molecule and erythrose, and in turn, the erythrose molecule can be transformed into glyceraldehyde (analysed in traces) by an oxidative attack. Finally formic acid, as the over-oxidation species, was obtained. The presence of glucaric acid has been also observed indicating the oxidation of both C1 and C6 atoms of glucose. The contemporaneous or successive oxidation of C1 and C6 can be explained by considering different glucose adsorption modes on the surface of the photocatalysts. The primary oxidation at the C6 of glucose to give glucuronic acid, reported in literature as one of the intermediates in catalytic oxidation of glucose,1 was not observed under the experimental condition used in this work, however in Scheme 1 it has been also hypothesized, and represented in brackets.

4. Conclusions

We can conclude that, although fructose, gluconic acid, arabinose, erythrose and formic acid were observed as oxidation products when bare TiO2 or HPA/TiO2 composite materials were used, depending on the photocatalyst used a different reaction extent and distribution of intermediate oxidation products were observed. Moreover, traces of glucaric acid and glyceraldehyde were also found in some runs.

The presence of PW12 or PW11 in the composite material influenced the reaction mechanism although no reactivity was observed in the presence of bare PW12, whereas bare PW11 induced only isomerization of the glucose.

The C mass balance was virtually accomplished in the presence of commercial TiO2 P25 and all of the HPA/TiO2 composites. The experiments carried out in the presence of samples containing TiO2 solvothermically prepared, instead, indicated a carbon unbalance, due to strong adsorption of reaction products on the photocatalyst surface.

References

  1. A. Corma, S. Iborra and A. Velty, Chem. Rev., 2007, 107, 2411 CrossRef CAS PubMed.
  2. Y. Lin and S. Tanaka, Appl. Microbiol. Biotechnol., 2006, 69, 627 CrossRef CAS PubMed.
  3. R. Geyer, P. Kraak, A. Pachulski and R. Schödel, Chem. Ing. Tech., 2012, 84, 513 CrossRef CAS PubMed.
  4. J. N. Chheda, Y. Roman-Leshkov and J. A. Dumesic, Green Chem., 2007, 9, 342 RSC.
  5. R. R. Davda, J. W. Shabaker, G. W. Huber, R. D. Cortright and J. A. Dumesic, Appl. Catal., B, 2005, 56, 171 CrossRef CAS PubMed.
  6. M. Comotti, C. D. Pina and M. Rossi, J. Mol. Catal. A: Chem., 2006, 251, 89 CrossRef CAS PubMed.
  7. T. Mallat and A. Baiker, Chem. Rev., 2004, 104, 3037 CrossRef CAS PubMed.
  8. S. Ramchandran, P. Fontanille, A. Pandey and C. Larroche, Food Technol. Biotechnol., 2006, 44, 185 Search PubMed.
  9. J. Z. Liu, L. P. Weng, Q. L. Zhang, H. Xu and L. N. Ji, Biochem. Eng. J., 2003, 14, 137 CrossRef CAS.
  10. M. Besson and P. Gallezot, Catal. Today, 2000, 57, 127 CrossRef CAS.
  11. C. Baatz, B. Thielecke and Y. Prüße, Appl. Catal., B, 2007, 70, 653 CrossRef CAS PubMed.
  12. S. Hermans and M. Devillers, Appl. Catal., A, 2002, 235, 253 CrossRef CAS.
  13. N. Thielecke, K. D. Vorlop and U. Prüße, Catal. Today, 2007, 122, 266 CrossRef CAS PubMed.
  14. M. Comotti, C. Della Pina, E. Falletta and M. Rossi, J. Catal., 2006, 244, 122 CrossRef CAS PubMed.
  15. P. Beltrame, M. Comotti, C. Della Pina and M. Rossi, Appl. Catal., A, 2006, 297, 1 CrossRef CAS PubMed.
  16. S. Biella, L. Prati and M. Rossi, J. Catal., 2002, 206, 242 CrossRef CAS.
  17. T. Ishida, N. Kinoshita, H. Okatsu, T. Akita, T. Takei and M. Haruta, Angew. Chem., 2008, 120, 9405 CrossRef PubMed.
  18. G. Palmisano, E. García-López, G. Marcì, V. Loddo, S. Yurdakal, V. Augugliaro and L. Palmisano, Chem. Commun., 2010, 46, 7074 RSC.
  19. A. Di Paola, E. García-López, G. Marcì and L. Palmisano, J. Hazard. Mater., 2012, 3–29, 211 Search PubMed.
  20. J. C. Colmenares, R. Luque, J. M. Campelo, F. Colmenares, Z. Karpiński and A. A. Romero, Materials, 2009, 2, 2228 CrossRef CAS PubMed.
  21. J. C. Colmenares, A. Magdziarz, K. Kurzydlowski, J. Grzonka, O. Chernyayeva and D. Lisovytskiy, Appl. Catal., B, 2013, 134–135, 136 CrossRef CAS PubMed.
  22. J. C. Colmenares, A. Magdziarz and A. Bielejewska, Bioresour. Technol., 2011, 102, 11254 CrossRef CAS PubMed.
  23. S. Yurdakal, G. Palmisano, V. Loddo, V. Augugliaro and L. Palmisano, J. Am. Chem. Soc., 2008, 130, 1568 CrossRef CAS PubMed.
  24. R. Chong, J. Li, Y. Ma, B. Zhang, H. Han and C. Li, J. Catal., 2014, 314, 101 CrossRef CAS PubMed.
  25. C. Streb, Dalton Trans., 2012, 1651 RSC.
  26. G. Marcì, E. García-López and L. Palmisano, Eur. J. Inorg. Chem., 2014, 1, 21 CrossRef PubMed.
  27. G. Marcì, E. García-López, L. Palmisano, D. Carriazo, C. Martin and V. Rives, Appl. Catal., B, 2009, 90, 497 CrossRef PubMed.
  28. G. Marcì, E. García-López and L. Palmisano, Appl. Catal., A, 2012, 421–422, 70 CrossRef PubMed.
  29. N. Mizuno and M. Misono, Chem. Rev., 1998, 98, 199 CrossRef CAS PubMed.
  30. I. V. Kozhevnikov, J. Mol. Catal. A: Chem., 2007, 262, 86 CrossRef CAS PubMed.
  31. N. Haraguchi, Y. Okaue, T. Isobe and Y. Matsuda, Inorg. Chem., 1994, 33, 1015 CrossRef CAS.
  32. M. T. Pope and A. Müller, Angew. Chem., Int. Ed. Engl., 1991, 30, 34 CrossRef PubMed.
  33. C. R. Rocchiccioli-Deltcheff, M. Fournier, R. Franck and R. Thouvenot, Inorg.Chem., 1983, 22, 207 CrossRef CAS.
  34. M. S. P. Francisco and V. R. Mastelaro, Chem. Mater., 2002, 14, 2514 CrossRef CAS.
  35. U. Lavrenčič Štangar, B. Orel, A. Régis and Ph. Colomban, J. Sol-Gel Sci. Technol., 1997, 8, 965 CrossRef.
  36. J. Tauc, Mater. Res. Bull., 1968, 3, 37 CrossRef CAS.
  37. F. Ma, T. Shi, J. Gao, L. Chen, W. Guo, Y. Guo and S. Wang, Colloids Surf., A, 2012, 401, 116 CrossRef CAS PubMed.
  38. Y. Roman-Leshkov, M. Moliner, J. A. Labinger and M. E. Davis, Angew. Chem., Int. Ed., 2010, 49, 8954 CrossRef CAS PubMed.

Footnote

Paper presented at FineCat 2015, Palermo, Italy, April 8-9, 2015.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.