Next Article in Journal
Electrospun Functional Materials toward Food Packaging Applications: A Review
Previous Article in Journal
Bimodal Nd-Doped LuVO4 Nanoprobes Functionalized with Polyacrilic Acid for X-Ray Computed Tomography and NIR Luminescent Imaging
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Combined AOPs for Formaldehyde Degradation Using Heterogeneous Nanostructured Catalysts

Department of Industrial Engineering, University of Padova, Via Marzolo 9, 35131 Padova, Italy
*
Authors to whom correspondence should be addressed.
Nanomaterials 2020, 10(1), 148; https://doi.org/10.3390/nano10010148
Submission received: 4 December 2019 / Revised: 28 December 2019 / Accepted: 9 January 2020 / Published: 14 January 2020
(This article belongs to the Section Energy and Catalysis)

Abstract

:
In this paper we studied the combination of advanced oxidation processes (AOPs), i.e., TiO2-based photocatalysis and photo-Fenton process, on the degradation of aqueous solutions containing a low (90 ppm) concentration of formaldehyde. Heterogeneous nanostructured catalysts, supported on polymeric nanofibers, were used; for comparison, some homogeneous or partly heterogeneous systems were also analyzed. Furthermore, to make the process more sustainable (in terms of costs and safety) no hydrogen peroxide was added to the system. The results showed that the combination of AOPs gave a synergy since the presence of iron was beneficial in promoting the photocatalytic activity of TiO2 while TiO2 was beneficial in promoting the photo-Fenton reaction. Moreover, very good results were obtained using fully heterogeneous nanostructured catalysts (based on TiO2 and FeSO4), without the need to add H2O2.

1. Introduction

Terrorism is a threat that, in recent decades, has become increasingly fearsome and destructive. It is known that chemical weapons and toxic compounds are used by terrorists in their attacks. For this reason, toxic industrial materials (TIMs), which are commonly used in the chemical industry, are the subject of discussion in many countries, where governments want to analyze and optimize the conditions of use and disposal of these hazardous chemicals. TIMs are characterized by high toxicity and wide availability [1]. This problem has led to the introduction of restrictive regulations on trade of these materials and on safety in chemical industries which, having a large quantity of toxic materials, are a sensitive target for possible terrorist actions and, therefore, require an adequate TIMs degradation system. To limit the spread of these substances, wastewater from industries, which could contain toxic compounds in high concentrations, should also be treated. One of the most dangerous TIMs is formaldehyde, a suspected carcinogenic chemical compound characterized by high toxicity, which is widely used in various sectors of the chemical industry (synthetic resins, insulators, paints, textiles) [2].
Advanced oxidation processes (AOPs) are considered the most effective treatments for the removal of these species from water [3], because they allow a complete mineralization even of the most refractory organic compounds, thus destroying wastewater pollutants and transforming them into less and even non-toxic products, thereby providing an ultimate solution for wastewater treatment.
Among the most widespread radical species used in AOPs, hydroxyl radical plays a primary role, since it is the most reactive oxidizing agent in water treatment (having a high oxidizing power [3]), it is very non-selective in its behavior and rapidly reacts with numerous species. Because hydroxyl radicals have a very short lifetime, they are only in situ produced during application through different methods, including a combination of oxidizing agents (e.g., H2O2), irradiation (such as ultraviolet (UV) light, or ultrasound), and catalysts (such as Fenton reagent, Fe2+, or titanium dioxide, TiO2) [4]. Indeed, hydroxyl radicals can be initiated by photons in the presence of catalysts (the most common being titanium dioxide, TiO2) or oxidants such as H2O2 (a H2O2 molecule is cleaved by UV irradiation to generate two OH•); they are also produced in the presence of some metals (e.g., iron) which are able to activate H2O2. For example, in the so-called Fenton process, H2O2 reacts with Fe2+ to generate strong reactive species, like hydroxyl radicals, although other substances such as ferryl ions are proposed [5].
Several papers have already reported on the degradation of aqueous formaldehyde solution by photocatalytic process using titanium dioxide [6,7,8,9]. Some of them have used homogeneous catalysis, i.e., TiO2 particles suspended in contaminated water, which makes it necessary to recover them after treatment while the heterogeneous (i.e., supported) catalyst configurations, on the other hand, eliminate the need for catalyst filtration but generally result in a significant reduction in system efficiency [10].
Fenton and photo-Fenton processes have already been proposed for formaldehyde removal from wastewater [11,12,13]. In general, the Fenton and photo-Fenton processes require the use of ferrous ions (Fe2+) and hydrogen peroxide (H2O2) [14]. Thus, large amounts of H2O2 need to be irreversibly consumed in real application to treat wastewater containing a high concentration of organic pollutants, and considerable iron dosage is indispensable if higher efficiency is required [15,16]. In the Fenton process, there is a need for continuous addition of ferrous ion into the reaction medium for the reaction to proceed further. This disadvantage could be overcome by the use of photo-Fenton reagent, which is a cyclic process and regenerates Fe2+ ion [15], thus limiting the deposition of ferric ion sludge.
Based on the similarity between the mechanism of destruction in the case of different advanced oxidation processes (AOPs) [17], some research efforts focusing on the synergism between several AOPs have already been published. Mokhbi et al. [18] studied the effectiveness of photocatalysis with Fenton’s reagent for treating oily wastewater; Kim et al. [19] studied the combined photocatalytic- and photo-Fenton system for treating some organic substrates (i.e., phenol, benzoic acid, and methanol); Xu et al. [15] studied interactions between TiO2 and Fe3+/H2O2 to degrade bisphenol A. They all used unsupported catalysts, i.e., TiO2 and Fe2+ were put in solution, in the presence of H2O2 and acidic pH, and found that synergistic effects resulted. In other works, the combination of TiO2 and Fe ions has been proposed but without exploiting Fenton or photo-Fenton reactions; the main aim of those works was, indeed, to dope titania in order to modify its photocatalytic features, such as visible light absorption and photocatalytic efficiency. For example, Khanmohammadi et al. [20] studied the use of unsupported catalyst, composed of sol-gel synthesized Fe2O3-TiO2 nano hybrid, in the degradation of liquid-phase formaldehyde in the presence of ultraviolet and visible irradiation. They found that the degradation took place according to a two-step mechanism: first, formaldehyde was converted to formic acid and then this was converted to H2O and CO2 during a photocatalytic reaction via radical mechanism in the presence of UV irradiation. In a similar way, Siddhapara et al. [21] used sol-gel TiO2 nanoparticles doped with a transition metal (Mn, Fe, Co) to degrade liquid formaldehyde solution; they found that the substitution with transition metals enhanced photocatalytic efficiency since transition metals increased defect sites and acted as a permanent space charge region, whose electric force improved the separating efficiency of electron-holes. Li et al. [22] studied the effect of the content of Fe on the photocatalytic performance of doped TiO2 on the degradation of gaseous formaldehyde and found that doped TiO2 exhibited higher photocatalytic activity under visible light irradiation. Cheng et al. [23] mixed TiO2 with iron oxide to produce modified photocatalysts (Fe/TiO2) and then coated them on a supporting medium of fiberglass and studied the degradation of gaseous formaldehyde using these supported catalysts. They concluded that the enhancement of formaldehyde decomposition by iron-doped TiO2 may be attributed to the facts that the doped iron ions could effectively enhance the absorption of UV-visible wavelength, increase the photocatalytic reactivity and retard the recombination of electron and electron-hole pairs, which thus contributed to the increase of formaldehyde decomposition efficiency. Chun et al. [24] studied the photocatalytic degradation of gaseous pollutants (benzene, toluene, ethyl benzene, and o-xylene (BTEX)) by preparing the Fe-TiO2 catalyst by the sol-gel technique and electrospinning it in presence of polyvinyl pyrrolidone; they found enhanced photocatalytic activity in presence of optimal Fe:TiO2 ratio.
However, authors using Fe-doped titania [20,21,22,23,24] did not mention the photo-Fenton reaction and did not adjust pH accordingly.
Nevertheless, in literature there are already some examples of combination of photocatalysis and photo-Fenton using supported catalysts. For example, Bouras et al. [25] combined a titania-supported photocatalyst as well as Photo-Fenton oxidation, by using FeCl2⋅4H2O in solution, for the decolorization of aqueous solution of the azo dye Basic Blue 41. Only very few examples using a fully supported catalytic system based on TiO2 and Fe ions and exploiting photo-Fenton reaction can be found. For example, Quici et al., degraded oxalic acid [26], while Shao et al. [27] used TiO2 and Fe(III) meso-tetraphenylporphyrin on polystyrene nanofibers to degrade methyl orange under visible light. They electrospun solutions simultaneously containing polystyrene, porphyrin and TiO2 and found that this system, in presence of added H2O2, coupling photocatalysis and Fenton reaction gave the best performance. Zhang et al. [28] used TiO2/Cu2O composite coated on glass matrix to degrade methylene blue using both photocatalysis and the Fenton process, using in situ generated H2O2. In the paper of Zhu et al. [29], beta-iron oxyhydroxide (β-FeOOH) nanostructures on electrospun TiO2 nanofibers were synthesized and used to degrade methyl orange. They showed that the degradation was faster in the presence of added H2O2 but it still took place also without adding it.
In this paper we study the combination of TiO2-based photocatalysis and photo-Fenton process on the degradation of aqueous solutions containing a low (90 ppm) concentration of formaldehyde to understand if synergy can be revealed. Heterogeneous nanostructured catalysts, supported on polymeric nanofibers, have been used; for comparison, some homogeneous or partly heterogeneous systems have also been analyzed. Furthermore, to make the process more sustainable (in terms of costs and safety) no hydrogen peroxide has been added to the system. To the best of our knowledge, this combination of advanced oxidation processes has never been studied in the literature for formaldehyde degradation.

2. Materials and Methods

2.1. Materials

Polyacrylonitrile (PAN, Mw = 150,000 g/mol, Sigma Aldrich, St. Louis, MO, USA) was used as support for the catalytic systems in electrospun form; N,N-dimethylformamide (DMF, Sigma-Aldrich, St. Louis, MO, USA) was used as electrospinning solvent. Ferrous sulfate (FeSO4⋅7H2O) was used as a catalyst for the photo-Fenton process and was supplied by Sigma-Aldrich (St. Louis, MO, USA). A crystalline, titanium dioxide powder was used for photocatalytic process, i.e., Aeroxide® TiO2 P90 (d = 14 nm, BET = 90 m2/g, 90/10 anatase/rutile, crystal size = 11 nm [30]), supplied by Evonik (Essen, Germany). We decided to use this one (instead of the most commonly used P25) because some preliminary trials (unpublished results) have shown that P90 was more active than P25 in our degradation systems. Some other authors reported similar results for different pollutants when using supported TiO2-based catalysts [30,31].
The aqueous formaldehyde solution for testing the catalytic activity was prepared by diluting formalin (37% w/w of formaldehyde and 10% w/w of methanol), supplied by Scharlau (Barcelona, Spain), with distilled water to reach a final formaldehyde concentration of 90 ppm. When the photo-Fenton process was considered, it was necessary to use 0.2 N sulfuric acid (H2SO4 at 98% w/w, Sigma Aldrich, St. Louis, MO, USA) to set the pH around 3, because this type of process is favored by acid pH [12]. Hydrogen peroxide was used as was a 30% (w/w) solution, supplied by Sigma Aldrich (St. Louis, MO, USA).
Finally, pentafluorobenzylhydroxylamine hydrochlorinated (PFBHA-HCl), purchased by Alfa Aesar (Haverhill, MA, USA), and 1,2-dibromopropane 98% w/w as internal standard (Sigma-Aldrich, St. Louis, MO, USA) dissolved into hexane (Sigma-Aldrich, St. Louis, MO, USA) have been used for the derivatization and extraction process, needed to quantify formaldehyde content in aqueous solutions by gas chromatography-mass spectrometry (GC-MS), according to the U.S. Environmental Protection Agency (EPA) method 556.

2.2. Preparation of the Nanostructured Catalytic Systems

Supported nanostructured catalysts have been used. The support was made by electrospun PAN nanofibers, while electrospraying was used for the catalysts’ deposition. In this way, the catalysts were completely placed on the upper surface of the membrane and so they were homogeneously irradiated by ultraviolet (UV) rays [32]. A schematic procedure is depicted in Figure 1. The use of nanofibers membranes as support allowed to have a very large surface area where the catalysts can be deposited.
The first step in preparation was the production of PAN nanofiber membranes by electrospinning, using a solution of PAN in DMF at 5% w/w, which must be stirred for not less than 15 h. The homogeneous mixture was then transferred into a 5 mL syringe fitted with a 27 G gauge needle. The syringe was put horizontally on a pump (NE-300 single syringe pump, NewEra Pump System Inc., Farmingdale, NY, USA) and the positive electrode of the high voltage power supply (Gamma High Voltage Inc., Ormond Beach, FL, USA), capable of generating voltages up to 60 kV, was clamped to the metal needle tip. The nanofibers were deposited on a metallic rotating collector. The process parameters are reported in Table 1.
For the catalysts deposition by electrospraying process, the catalysts were suspended into ethanol. TiO2 was used at 5% w/w; this concentration was chosen in order to obtain a catalyst content of about 0.14 mg/cm2, which was, based on our previous experience [32], the optimal one to have a widespread distribution of titanium dioxide, avoiding great clusters, agglomeration and clogging of the nanostructured substrate. Moreover, at higher concentrations, an increase in opacity and light scattering of TiO2 particles occurs, leading to a decreased passage of irradiation through the sample [10]. For ferrous sulfate, based on mass balance calculation, a 4% w/w solution in ethanol should be enough to have the same catalyst content of titania and similar to that used in literature for a homogenous photo-Fenton system [11]. However, due to a partial loss of FeSO4 experienced during the electrospray process, owing probably to an interaction of the Fe2+ with the electric field, a 10% w/w FeSO4 solution in ethanol has to be used. When the two catalysts were employed together, TiO2 was used at 5% w/w and the FeSO4 at 10% w/w dispersion in ethanol, in order to deposit a 50:50 TiO2:FeSO4 catalytic system on to the PAN support.
The catalysts’ dispersions (TiO2 and/or FeSO4 in ethanol) were sonicated for 40 min at 40% of amplitude and constant temperature by using ultrasonic probe VC505® of Sonics & Materials (power of 500 W and a probe length of 254 mm, Newtown, CT, USA). Then, Dynasylan® 4144 was added at 1% w/w of ethanol, followed by a further sonication (15 min).
Similarly, to the electrospinning, the catalysts’ dispersion was put in a syringe of 5 mL volume and then it was electrosprayed onto the preformed electrospun PAN membrane thanks to a voltage generated between the metallic needle and the collector. The whole production process was carried out at room temperature and the humidity was kept constant by flowing dried compressed air.
In the following, membranes are named PAN_TiO2, PAN_Fe and PAN_TiO2_Fe if the catalyst deposited was TiO2, FeSO4 or TiO2+FeSO4, respectively. In this work, we referred to homogeneous process when the catalyst (Fe2+) and/or oxidant (H2O2) were in the same (liquid) phase of formaldehyde (i.e., they are simply added to formaldehyde solution) while heterogeneous one referred to the use of supported TiO2 and/or Fe catalyst prepared by electro-spinning and -spraying.
Concentrations of catalysts, in both homogeneous and heterogeneous experiments, have been fixed based on optimal compositions reported in literature [11] or based on our previous experience [32]: for homogeneous catalysis, FeSO4 and H2O2 were used at 24 and 400 ppm, respectively, while in the heterogeneous supported system around 0.15 mg/cm2 of catalyst has been deposited (the exact values are obtained by thermogravimetric analysis (TGA) and are reported in the following).

2.3. Testing of the Catalytic Systems in the Degradation of Aqueous Formaldehyde Solution

The set-up used for the experiments is shown in Figure 2. It consisted mainly of a Petri dish with a diameter of 15 cm and a height of 2 cm, in which 70 mL of aqueous formaldehyde solution (90 ppm) is placed; a magnetic stirrer provided a good mixing inside the system. The catalytic membrane, comprising titanium dioxide and/or ferrous sulfate, with a 50 cm2 area, was placed on the bottom of the Petri dish and it was kept fixed using eight stainless steel weights. UV irradiation is provided by a lamp (16 W UV Stylo E16, Light Progress, Anghiari, Italy) with a spectral peak centered around 250 nm. The distance between the lamp and the membrane was constant (2 cm). For some trials, ferrous sulfate was added directly to the aqueous solution. In this case the experiment started after the complete solubilization of the salt into the solution.
The catalytic process lasted 4 h. During the tests, some samples were taken from the aqueous solution. These samples were subjected to a derivatization process according to EPA method 556, before being injected into a GC-MS to quantify formaldehyde content.
It is known that bubbling air within the formaldehyde solution gives some benefits in terms of degradation rate [6] through the electron transfer reaction generating superoxide radical anion:
e + O 2     O 2
However, we preferred to avoid the use of air in order to avoid misleading results due to formaldehyde stripping by the bubbling air. Nonetheless, due to traces of (electron acceptor) impurities or adventitious O2 in the solution (since the solution is in direct contact with ambient air) some superoxide radical anions can still form [6].

2.4. Characterization

All the membranes are characterized in terms of morphology by scanning electron microscopy (SEM, Obducat CamScan MX2500, Cambridge, UK) and transmission electron microscopy (TEM, FEI model Tecnai G12, Hillsboro, OR, USA). Sample preparation for TEM imaging have been carried out as follows: the membrane samples were embedded in a proper epoxy resin and slices of 100 nm thickness were obtained by ultramicrotomy of the cured resin.
Furthermore, catalytic content deposited onto the PAN support was investigated through TGA (SDT Q600, TA Instruments, New Castle, DE, USA); the analysis was performed using alumina pan with air flow of 100 cc/min and a heating rate of 20 °C/min, from room temperature to 1000 °C.
The quantification of formaldehyde was carried out by withdrawing samples from the aqueous solution during the degradation tests and derivatizing them using o-(2,3,4,5,6-Pentafluorobenzyl)hydroxylamine hydrochloride (PFBHA-HCl). The oxime derivative formed was extracted from water with 4 mL hexane. The extract was processed through an acidic wash step and then analyzed by GC-MS (Gas Chromatograph Trace 1300® coupled with the Single Quadrupole Mass Spectrometer ISQ QD®, both purchased from Thermo Scientific, Waltham, MA, USA). The GC column was a non-polar DB5 capillary column (0.25 mm i.d., 30 m length, supplied by Agilent, Santa Clara, CA, USA). According to the EPA method 556, an internal standard (1,2-dibromopropane) has been used to obtain quantitative data. A calibration curve has also been built before starting the experimental tests, by analyzing formaldehyde solutions of known concentration. The target analytes (oxime derivative, to quantify formaldehyde, and internal standard) were identified (m/z 225 and 121, respectively) and quantified by using a calibration curve. Each data point has been obtained as the mean value of three replicates.

3. Results and Discussion

3.1. Catalytic Membranes Characterization

All the produced membranes were characterized by SEM analysis in order to gain a better understanding of the morphology. The SEM micrographs of the PAN and PAN_TiO2 membranes are reported in Figure 3. It can be observed that PAN nanofibers were smooth, with a mean diameter of ~300 nm. The TiO2 nanoparticles had the tendency to create clusters when spread over the nanofibers surface and some large aggregates were formed. However, the catalytic particles had an almost homogeneous distribution upon PAN support without clogging it.
TEM images of PAN-TiO2 are reported in Figure 4. In Figure 4a, the cross sections of the nanofibers can be observed; in Figure 4b, TiO2 nanoparticles can be seen around the nanofibers (the cross section of the fiber is not circular due to cutting direction); TiO2 clusters were, mainly, less than 500 nm. From SEM and TEM analysis of PAN_Fe membranes (Figure 5), it was evident that this kind of catalyst also covered almost completely the nanofibers of PAN; in contrast to what happens for TiO2, the clusters had larger dimension and no spherical shape. The difference between the two catalysts is more clear if the TEM image is analysed: ferrous sulfate had bigger dimensions and irregular shape, while TiO2 appeared in spherical clusters with submicron mean diameter.
The catalyst content of each prepared membrane was measured by TGA analysis, analyzing the residue at high temperature (1000 °C), taking into account that TiO2 did not lose any weight at that temperature while FeSO4 lost 70% of its initial weight. Based on TGA results, the specific catalyst content for each membrane has been measured; the results are reported in Table 2.

3.2. Photocatalytic Degradation of the Formaldehyde

The photocatalytic degradation of formaldehyde, in the presence of TiO2, takes place according to the following reactions (Scheme 1, [33]):
The results for catalytic degradation with PAN_TiO2 membrane are reported in Figure 6; PAN_TiO2 membrane contained 0.14 mg/cm2 of TiO2 (i.e., optimal composition defined elsewhere [32]) which corresponded to a total amount of TiO2 of 7 mg in the system.
For heterogeneous catalytic system PAN_TiO2, an increase in the formaldehyde content was observed during the first stage of degradation. This can be rationalized considering that the degradation of methanol (which was contained at 10 wt% in the formalin solution as stabilizer to limit formaldehyde polymerization) led to the formation of formaldehyde, according to the following reactions (Scheme 2, [34]):
The degradation of the methanol can start as soon as a hydroxyl radical is produced and it may compete with formaldehyde to consume hydroxyl radicals resulting in the retardation of the oxidation reaction of formaldehyde. Both methanol and formaldehyde formed formic acid during their degradation. Indeed, the initial pH of the solution was around 6.5–7, and it was not adjusted, owing to the pH-independent photocatalytic activity of TiO2 [19], but during the test, it decreased to 4.2 and then rose to 5, as already reported in similar studies [6,7], thus implying the formation of formic acid. Therefore, we can assess that, since both methanol and formaldehyde were oxidized to formic acid (and then to CO2 and water), these compounds were responsible for the drop in the pH of the solution. As CO2 was liberated from the solution, pH increased [7].
However, it was evident that after 4 h the abatement of the formaldehyde content was still low.
The addition of H2O2 (400 ppm, according to optimal content reported in [11]) in aqueous solution to the PAN-TiO2 system (PAN_TiO2 + H2O2 in Figure 6) resulted in greater oxidation with respect to neat PAN-TiO2 and this was likely attributed to the enhanced charge separation induced by H2O2 as an electron acceptor and the reductive conversion of H2O2 to OH• [7]. Moreover, the methanol degradation was very fast and the increase in the formaldehyde content was not revealed (since the first sampling point was after 1 h).
Comparing homogeneous system using H2O2 only (400 ppm, without catalytic membrane, curve H2O2 in Figure 6) with the previous heterogeneous one (PAN_TiO2 + H2O2), it can be noted that the higher removal rate was mainly due, especially in the first 2 h, to the direct oxidation of H2O2 onto the organic substrate. The trend of the degradation of formaldehyde with only H2O2 was very similar to that reported by Guimarães et al. [7]. Since also in H2O2-system the pH decreased from 6.5 to 3.8, the formation of formic acid can still be speculated. After the H2O2 was completely consumed, the degradation reaction stopped and the CH2O concentration reached a plateau while in the presence of TiO2 it seemed to further decrease.

3.3. Photo-Fenton Degradation of Formaldehyde

Photo-Fenton degradation of formaldehyde was also studied. Several catalytic, homogeneous and heterogeneous systems have been compared, to gain deep understanding of the process. In this work, we referred to homogeneous process when the catalyst (Fe2+) and/or oxidant (H2O2) were in the same (liquid) phase of formaldehyde while the heterogeneous one referred to the use of supported Fe catalyst.
Generally speaking, photo-Fenton systems are based on the contemporary use of Fe ions and hydrogen peroxide: ferrous ions react with hydrogen peroxide under the irradiation of UV light, producing hydroxyl radicals with powerful oxidizing abilities. While in the Fenton reaction Fe3+ ions accumulate in the system and the reaction stops once all Fe2+ ions are consumed, in the photo-Fenton reaction the regeneration of ferrous ions by photo-reduction of ferric ions occurs according to Scheme 3:
The newly generated ferrous ions react again with H2O2 generating hydroxyl radical and ferric ion and, in this way, the cycle continues. Therefore, the rate of degradation of organic pollutants in photo-Fenton reaction increases with respect to the Fenton one [35]. The photo-Fenton process offers better performance at pH 3.0, when the hydroxyl-Fe3+ complexes are more soluble and Fe(OH)2+ is more photoactive [36]. However, although ferrous ions are regenerated, generally, a large amount of H2O2 needs to be irreversibly consumed.
Results for formaldehyde degradation in a homogeneous photo-Fenton process, with and without added H2O2 (Fe/H2O2 and Fe curves, respectively) as well as heterogeneous one (PAN_Fe curve) are reported in Figure 7; homogeneous photolytic oxidation (H2O2 curve in Figure 7) is also reported for comparison. In photo-Fenton processes (i.e., whenever Fe was present) the pH has been adjusted to 3 using 0.2 N sulfuric acid. When ferrous sulfate and H2O2 were used in solution, their concentrations were 24 and 400 ppm, respectively, i.e., set at the optimal conditions identified elsewhere [11]. The heterogeneous PAN-Fe catalyst comprised 0.12 mg/cm2 of FeSO4 (i.e., similar content to the TiO2 content of the previous experiment, i.e., optimal composition defined elsewhere [32]), which corresponded to a total of 6 mg of ferrous sulfate in the system. All the trials have been carried out under UV irradiation; it shall be kept in mind that in presence of iron sulfate the UV absorbance of the system around 290 nm was quite high, owing to the absorption properties of this compound [23].
Within the homogeneous processes (curves Fe, H2O2 and Fe/H2O2 in Figure 7), the photo-Fenton Fe/H2O2 demonstrated the best performance; however, also the use of iron compound alone (Fe curve), without adding H2O2, gave good results since the final formaldehyde concentration was the same as for Fe/H2O2, although the initial rate was lower. Similar results have already been reported by other authors [29,37] for different compounds, i.e., the addition of H2O2 to Fe catalyst was beneficial, but the degradation took place also without adding H2O2.
There can be several reasons for this. First, it should also be kept in mind that small amount of H2O2 can also be developed during UV irradiation of formaldehyde [9]. Secondly, ferrous ion can react with dissolved oxygen in water (Scheme 4) forming reactive oxidants capable of oxidative degradation. Indeed, superoxide radical anions can form, by reaction of Fe2+ with oxygen of the air dissolved into the solution [6], leading to the formation of H2O2 [38]; moreover, Fe2+ in the UV light conditions, can be partially transformed into Fe3+, the Fe3+ in solution can be hydrolyzed to hydroxylable Fe(OH)2+, Fe (OH)2+ also can be transformed into Fe2+ under ultraviolet light, while producing OH radicals to form a Fe3+/Fe2+ cycle of reactions [11,39,40]:
Therefore, in the presence of Fe2+ and UV light several reactions can take place to generate oxidizing species (OH•, O 2 and H2O2) without the need to add external hydrogen peroxide.
Lastly, it could also be possible that the formic acid, deriving from degradation of formaldehyde, created a complex compound with iron [41] which could be photolyzed and even photocatalyzed in the reaction medium, generating Fe2+, H2O2 and other active radical species [38], in a similar way to that reported for oxalic acid by Quici et al. [26].
The effectiveness of Fe catalyst also without H2O2 is very important since avoiding the use of H2O2 makes the process more economic and safer.
Also, a heterogeneous photo-Fenton process, i.e., using a supported Fe catalyst (PAN_Fe), has been shown to be active in formaldehyde degradation; however, as expected, the reaction rate was slower with respect to homogenous systems. This PAN_Fe system reached the same final formaldehyde content as photolytic oxidation (H2O2 curve) process (but without the use of costly and hazardous reagents like water peroxide) and only slightly higher concentration than homogenous photo-Fenton processes.

3.4. Photo-Catalytic-Fenton Degradation of Formaldehyde

After the separate analysis of photocatalytic and photo-Fenton processes, the combination of such AOPs has been studied.
Figure 8 reports the results when using a heterogeneous system based on photo-Fenton (PAN_Fe), photocatalysis (PAN_TiO2) and their combination (PAN_TiO2_Fe), under UV illumination, pH = 3, without the use of any added H2O2. It can be noted that the combination of TiO2 and Fe2+ showed the best result, i.e., there was a synergy deriving from the combination of photocatalysis and photo-Fenton since the combined process gave better results than the simple sum of the two distinct processes.
This synergy can be due to the fact that in presence of iron sulfate the UV absorbance of the system around 290 nm increased, owing to the absorption properties of this compound, therefore favoring UV activation [23]. Moreover, this synergy can be related to the roles of iron as an electron acceptor, to facilitate charge separation in TiO2 photocatalyst, thereby reducing the non-desired electron/hole recombination and thus increasing the rate of OH• formation through the first and second reaction in Scheme 1. Similar results have been reported for homogeneous TiO2-Fe system in phenol degradation [19], for Azo dye [25] and for bisphenol A [15] and for Fe doped TiO2 [20,21,23].
It was also reported that in this system TiO2 is beneficial to the fast conversion of Fe3+→Fe2+ [15], thus favoring the closing the Fe3+/Fe2+ cycle of reactions mentioned before when dealing with photo-Fenton. Also, the accumulated electrons in the conduction band of TiO2 are transferred to oxygen (coming from the air in contact with formaldehyde solution) on the TiO2 surface for the formation of O2− or O22−, which combines with H+ to form H2O2 [28]. The evolved H2O2 with FeSO4 formed Fenton reagent which can further degrade formaldehyde [28].
Summarizing, the presence of iron was beneficial in promoting the photocatalytic activity of TiO2 while TiO2 was beneficial in promoting the photo-Fenton reaction.
If we compare the results obtained with a fully heterogeneous system (PAN_TiO2_Fe, where Fe is deposited onto the support) and partly homogeneous one (PAN_TiO2+Fe, where 24 ppm of ferrous sulfate was in the aqueous solution, as before in photo-Fenton trials) (Figure 9), it can be noted that the degradation rate was slightly faster for the latter, probably owing to a higher photo-Fenton reaction rate; however, both systems reached a final CH2O concentration of around 10 ppm.
However, if we compare the results on a catalyst weight basis, i.e., considering the moles of formaldehyde degraded per mg of catalyst, it can be shown (Figure 10) that up to 3 h there were no significant differences for PAN_Fe and PAN_TiO2_Fe; however, over a longer period the best performing system was PAN_Fe. Obviously, this was an initial work and further studies on the optimal iron content need to be carried out.

4. Conclusions

Photocatalytic and photo-Fenton processes as well as their combination were studied to degrade formaldehyde aqueous solution; both homogeneous and heterogeneous catalysts were used.
The results showed that the combination of AOPs gave a synergy since the presence of iron was beneficial in promoting the photocatalytic activity of TiO2 while TiO2 was beneficial in promoting the photo-Fenton reaction.
Moreover, very good results have been obtained using fully heterogeneous nanostructured catalysts (based on TiO2 and FeSO4) without the need to add H2O2. In this way, a more economic formaldehyde treatment process has been developed.

Author Contributions

Conceptualization, R.B., M.R., A.M. and A.L.; Data curation, R.B. and C.B.; Funding acquisition, A.L.; Investigation, C.B. and L.C.; Methodology, R.B., M.R., A.M. and A.L.; Project administration, A.L.; Supervision, R.B.; Visualization, R.B.; Writing—original draft, R.B., C.B. and L.C.; Writing—review and editing, M.R., A.M., M.M. and A.L. All authors have read and agreed to the published version of the manuscript.

Funding

The research leading to these results has received funding from University of Padova, Department of Industrial Engineering, Twinning Project 2017.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bennett, M. TICs, TIMs, and Terrorists. Today’s Chem. Work 2003, 12, 21–25. [Google Scholar]
  2. Raja Priya, K.; Sandhya, S.; Swaminathan, K. Kinetic analysis of treatment of formaldehyde containing wastewater in UAFB reactor. Chem. Eng. J. 2009, 148, 212–216. [Google Scholar] [CrossRef]
  3. Deng, Y.; Zhao, R. Advanced Oxidation Processes (AOPs) in Wastewater Treatment. Curr. Pollut. Rep. 2015, 1, 167–176. [Google Scholar] [CrossRef] [Green Version]
  4. Dong, C.P.H.C.; Tang, Z. Advanced chemical oxidation: Its present role and potential future in hazardous waste treatment. Waste Manag. 1993, 13, 361–377. [Google Scholar]
  5. Pignatello, J.J.; Oliveros, E.; MacKay, A. Advanced oxidation processes for organic contaminant destruction based on the fenton reaction and related chemistry. Crit. Rev. Environ. Sci. Technol. 2006, 36, 1–84. [Google Scholar] [CrossRef]
  6. Shin, E.M.; Senthurchelvan, R.; Munoz, J.; Basak, S.; Rajeshwar, K.; Benglas-Smith, G.; Howell, B.C. Photolytic and photocatalytic destruction of formaldehyde in aqueous media. J. Electrochem. Soc. 1996, 143, 1562–1570. [Google Scholar] [CrossRef]
  7. Guimarães, J.R.; Turato Farah, C.R.; Maniero, M.G.; Fadini, P.S. Degradation of formaldehyde by advanced oxidation processes. J. Environ. Manag. 2012, 107, 96–101. [Google Scholar] [CrossRef]
  8. Sifang, L.; Guoliang, Y.; Guoqin, C. Low-temperature preparation and characterization of nanocrystalline anatase TiO2. J. Phys. Chem. C 2009, 113, 4031–4037. [Google Scholar]
  9. Shiraishi, F.; Nakasako, T.; Hua, Z. Formation of Hydrogen Peroxide in Photocatalytic Reactions. J. Phys. Chem. A 2003, 107, 11072–11081. [Google Scholar] [CrossRef]
  10. Kabra, K.; Chaudhary, R.; Sawhney, R.L. Treatment of hazardous organic and inorganic compounds through aqueous-phase photocatalysis: A review. Ind. Eng. Chem. Res. 2004, 43, 7683–7696. [Google Scholar] [CrossRef]
  11. Liang, J.; Liu, X.; Zhang, Z.; Wang, Y. Kinetics and Reaction Mechanism for Formaldehyde Wastewater Using UV-Fenton Oxidation. In Proceedings of the 2010 4th International Conference on Bioinformatics and Biomedical Engineering (CBBE), Chengdu, China, 18–20 June 2010; IEEE: Piscataway, NJ, USA, 2010; pp. 1–5. [Google Scholar]
  12. Kajitvichyanukul, P.; Lu, M.C.; Jamroensan, A. Formaldehyde degradation in the presence of methanol by photo-Fenton process. J. Environ. Manag. 2008, 86, 545–553. [Google Scholar] [CrossRef] [PubMed]
  13. Pang, H.C.; Fu, H.Y.; Gao, T.Z. Treatment of high concentration wastewater containing phenols and aldehyde from a small Phenolic resin plant. Adv. Mater. Res. 2014, 864–867, 1552–1555. [Google Scholar] [CrossRef]
  14. Lazar, M.A.; Varghese, S.; Nair, S.S. Photocatalytic Water Treatment by Titanium Dioxide: Recent Updates. Catalysts 2012, 2, 572–601. [Google Scholar] [CrossRef] [Green Version]
  15. Xu, L.; Meng, L.; Zhang, X.; Mei, X.; Guo, X.; Li, W.; Wang, P.; Gan, L. Promoting Fe3+/Fe2+ cycling under visible light by synergistic interactions between P25 and small amount of Fenton reagents. J. Hazard. Mater. 2019, 379, 120795. [Google Scholar] [CrossRef] [PubMed]
  16. Gogate, P.R.; Pandit, A.B. A review of imperative technologies for wastewater treatment II: Hybrid methods. Adv. Environ. Res. 2004, 8, 553–597. [Google Scholar] [CrossRef]
  17. Adewu-Yi, Y.G. Sonochemistry in environmental remediation. 2. Heterogeneous sonophotocatalytic oxidation processes for the treatment of pollutants in water. Environ. Sci. Technol. 2005, 39, 8557–8570. [Google Scholar] [CrossRef]
  18. Mokhbi, Y.; Korichi, M.; Akchiche, Z. Combined photocatalytic and Fenton oxidation for oily wastewater treatment. Appl. Water Sci. 2019, 9, 35. [Google Scholar] [CrossRef] [Green Version]
  19. Kim, H.E.; Lee, J.; Lee, H.; Lee, C. Synergistic effects of TiO2 photocatalysis in combination with Fenton-like reactions on oxidation of organic compounds at circumneutral pH. Appl. Catal. B Environ. 2012, 115–116, 219–224. [Google Scholar] [CrossRef]
  20. Khanmohammadi, M.; Bagheri, A.; Elmizadeh, H. Spectrophotometric evaluation of the photocatalytic degradation of formaldehyde by Fe2O3–TiO2 nano hybrid. J. Ind. Eng. Chem. 2014, 20, 1841–1844. [Google Scholar] [CrossRef]
  21. Siddhapara, K.; Shah, D. Study of photocatalytic activity and magnetic properties of Co, Mn metal ions doped nanocrystalline TiO2 prepared by Sol–Gel method. J. Cryst. Growth 2016, 452, 158–161. [Google Scholar] [CrossRef]
  22. Li, J.; Ren, D.; Wu, Z.; Xu, J.; Bao, Y.; He, S.; Chen, Y. Flame retardant and visible light-activated Fe-doped TiO2 thin films anchored to wood surfaces for the photocatalytic degradation of gaseous formaldehyde. J. Colloid Interface Sci. 2018, 530, 78–87. [Google Scholar] [CrossRef] [PubMed]
  23. Cheng, S.W.; Li, Y.H.; Yuan, C.S.; Tsai, P.Y.; Shen, H.Z.; Hung, C.H. An innovative advanced oxidation technology for effective decomposition of formaldehyde by combining iron modified nano-TiO2 (Fe/TiO2) photocatalytic degradation with ozone oxidation. Aerosol Air Qual. Res. 2018, 18, 3220–3233. [Google Scholar] [CrossRef]
  24. Chun, H.H.; Lee, J.Y.; Jo, W.K. Photocatalysis of low-concentration gaseous organic pollutants over electrospun iron-doped titanium dioxide nanofibers. Solid State Sci. 2013, 25, 103–109. [Google Scholar] [CrossRef]
  25. Bouras, P.; Lianos, P. Synergy effect in the combined photodegradation of an azo dye by titanium dioxide photocatalysis and photo-fenton oxidation. Catal. Lett. 2008, 123, 220–225. [Google Scholar] [CrossRef]
  26. Quici, N.; Morgada, M.E.; Piperata, G.; Babay, P.; Gettar, R.T.; Litter, M.I. Oxalic acid destruction at high concentrations by combined heterogeneous photocatalysis and photo-Fenton processes. Catal. Today 2005, 101, 253–260. [Google Scholar] [CrossRef]
  27. Shao, L.; Hu, B.; Dong, P.; Ji, W.; Qi, C. Electrospinning Fe(III)porphyrin/TiO2/poly(styrene) mixture:formation of a novel nanofiber photocatalyst for the photodegradation of methyl orange. J. Porphyr. Phthalocyanines 2010, 14, 993–999. [Google Scholar] [CrossRef]
  28. Zhang, Y.G.; Ma, L.L.; Li, J.L.; Yu, Y. In situ Fenton reagent generated from TiO2/Cu2O composite film: A new way to utilize TiO2 under visible light irradiation. Environ. Sci. Technol. 2007, 41, 6264–6269. [Google Scholar] [CrossRef]
  29. Zhu, T.; Ong, W.L.; Zhu, L.; Ho, G.W. TiO2 fibers supported β-FeOOH nanostructures as efficient visible light photocatalyst and room temperature sensor. Sci. Rep. 2015, 5, 10601. [Google Scholar] [CrossRef] [Green Version]
  30. Keulemans, M.; Verbruggen, S.W.; Hauchecorne, B.; Martens, J.A.; Lenaerts, S. Activity versus selectivity in photocatalysis: Morphological or electronic properties tipping the scale. J. Catal. 2016, 344, 221–228. [Google Scholar] [CrossRef]
  31. Kete, M.; Pavlica, E.; Fresno, F.; Bratina, G.; Štangar, U.L. Highly active photocatalytic coatings prepared by a low-temperature method. Environ. Sci. Pollut. Res. 2014, 21, 11238–11249. [Google Scholar] [CrossRef]
  32. Modesti, M.; Roso, M.; Boaretti, C.; Besco, S.; Hrelja, D.; Sgarbossa, P.; Lorenzetti, A. Preparation of smart nano-engineered electrospun membranes for methanol gas-phase photoxidation. Appl. Catal. B Environ. 2014, 144, 216–222. [Google Scholar] [CrossRef]
  33. Yang, J.; Li, D.; Zhang, Z.; Li, Q.; Wang, H. A study of the photocatalytic oxidation of formaldehyde on Pt/Fe2O3/TiO2. J. Photochem. Photobiol. A Chem. 2000, 137, 197–202. [Google Scholar] [CrossRef]
  34. Arana, J.; Nieto Martinez, J.A.; Herrera Melian, J.A.; Dona Rodrıguez, J.M.; Gonzalez Dıaz, O.; Perez Pena, J.; Bergasa, O.; Alvarez, C.; Mendez, J. Photocatalytic degradation of formaldehyde containing wastewater from veterinarian laboratories endez. Chemosphere 2004, 55, 893–904. [Google Scholar] [CrossRef] [PubMed]
  35. Kim, S.M.; Vogelpohl, A. Degradation of Organic Pollutants by the Photo-Fenton-Process. Chem. Eng. Technol. 1998, 21, 187–191. [Google Scholar] [CrossRef]
  36. Ameta, R.; Chohadia, A.K.; Jain, A.; Punjabi, P.B. Fenton and Photo-Fenton Processes. In Advanced Oxidation Processes for Wastewater Treatment: Emerging Green Chemical Technology, 1st ed.; Ameta, S., Ameta, R., Eds.; Academic Press: Cambridge, MA, USA, 2018; pp. 49–87. ISBN 9780128105252. [Google Scholar]
  37. Ghaly, M.Y.; Härtel, G.; Mayer, R.; Haseneder, R. Photochemical oxidation of p -chlorophenol by UV/H2O2 and photo-Fenton process. A comparative study. Waste Manag. 2001, 21, 41–47. [Google Scholar] [CrossRef]
  38. Augugliaro, V.; Litter, M.; Palmisano, L.; Soria, J. The combination of heterogeneous photocatalysis with chemical and physical operations: A tool for improving the photoprocess performance. J. Photochem. Photobiol. C Photochem. Rev. 2006, 7, 127–144. [Google Scholar] [CrossRef] [Green Version]
  39. Lee, C.; Keenan, C.R.; Sedlak, D.L. Polyoxometalate-Enhanced Oxidation of Organic Compounds by Nanoparticulate Zero-Valent Iron and Ferrous Ion in the Presence of Oxygen. Environ. Sci. Technol. 2008, 42, 4921–4926. [Google Scholar] [CrossRef] [Green Version]
  40. Benkelberg, H.; Warneck, P.; Chemie, M. Photodecomposition of Iron (III) Hydroxo and Sulfato Complexes in Aqueous Solution: Wavelength Dependence of OH and SO4-Quantum Yields. J. Phys. Chem. 1995, 99, 5214–5221. [Google Scholar] [CrossRef]
  41. Kowalik, P. Chemical pretreatment of formaldehyde wastewater by selected Advanced Oxidation Processes (AOPs). Chall. Mod. Technol. 2011, 2, 42–48. [Google Scholar]
Figure 1. Schematic procedure for the preparation of nanostructured catalytic systems.
Figure 1. Schematic procedure for the preparation of nanostructured catalytic systems.
Nanomaterials 10 00148 g001
Figure 2. Experimental set-up for testing catalytic activity in formaldehyde degradation.
Figure 2. Experimental set-up for testing catalytic activity in formaldehyde degradation.
Nanomaterials 10 00148 g002
Figure 3. Scanning electron microscope (SEM) micrographs of (a) PAN, (b,c) PAN_TiO2 membranes.
Figure 3. Scanning electron microscope (SEM) micrographs of (a) PAN, (b,c) PAN_TiO2 membranes.
Nanomaterials 10 00148 g003
Figure 4. Transmission electron microscope (TEM) images at different magnifications (a,b) of PAN_TiO2 membrane.
Figure 4. Transmission electron microscope (TEM) images at different magnifications (a,b) of PAN_TiO2 membrane.
Nanomaterials 10 00148 g004
Figure 5. SEM images of (a) PAN_Fe and (b) PAN_TiO2_Fe membranes; (c) TEM image of PAN_TiO2_Fe membrane.
Figure 5. SEM images of (a) PAN_Fe and (b) PAN_TiO2_Fe membranes; (c) TEM image of PAN_TiO2_Fe membrane.
Nanomaterials 10 00148 g005
Scheme 1. Mechanism of photocatalytic degradation of formaldehyde in aqueous solution.
Scheme 1. Mechanism of photocatalytic degradation of formaldehyde in aqueous solution.
Nanomaterials 10 00148 sch001
Figure 6. Formaldehyde concentration during photocatalytic experiment with and without added H2O2 (400 ppm). The maximum standard deviation of the experimental points was 1.9 ppm.
Figure 6. Formaldehyde concentration during photocatalytic experiment with and without added H2O2 (400 ppm). The maximum standard deviation of the experimental points was 1.9 ppm.
Nanomaterials 10 00148 g006
Scheme 2. Methanol photocatalytic degradation path.
Scheme 2. Methanol photocatalytic degradation path.
Nanomaterials 10 00148 sch002
Scheme 3. Photo-Fenton reactions.
Scheme 3. Photo-Fenton reactions.
Nanomaterials 10 00148 sch003
Figure 7. Formaldehyde concentration during photolysis and photo-Fenton experiment. The maximum standard deviation of the experimental points was 2.5 ppm.
Figure 7. Formaldehyde concentration during photolysis and photo-Fenton experiment. The maximum standard deviation of the experimental points was 2.5 ppm.
Nanomaterials 10 00148 g007
Scheme 4. Reactions between ferrous ion and dissolved oxygen.
Scheme 4. Reactions between ferrous ion and dissolved oxygen.
Nanomaterials 10 00148 sch004
Figure 8. Formaldehyde concentration profile using heterogeneous photo-Fenton (PAN_Fe), photocatalytic (PAN_TiO2) and their combination (PAN_TiO2_Fe) catalysts. The maximum standard deviation of the experimental points was 1.9 ppm.
Figure 8. Formaldehyde concentration profile using heterogeneous photo-Fenton (PAN_Fe), photocatalytic (PAN_TiO2) and their combination (PAN_TiO2_Fe) catalysts. The maximum standard deviation of the experimental points was 1.9 ppm.
Nanomaterials 10 00148 g008
Figure 9. Formaldehyde concentration profile using heterogeneous photo-catalytic-Fenton with Fe supported (PAN_TiO2_Fe) or dispersed in solution (PAN_TiO2+Fe). The maximum standard deviation of the experimental points was 1.8 ppm.
Figure 9. Formaldehyde concentration profile using heterogeneous photo-catalytic-Fenton with Fe supported (PAN_TiO2_Fe) or dispersed in solution (PAN_TiO2+Fe). The maximum standard deviation of the experimental points was 1.8 ppm.
Nanomaterials 10 00148 g009
Figure 10. Formaldehyde degraded per mg of catalyst for heterogeneous photo-Fenton (PAN_Fe), photocatalytic (PAN_TiO2) and their combination (PAN_TiO2_Fe) catalysts.
Figure 10. Formaldehyde degraded per mg of catalyst for heterogeneous photo-Fenton (PAN_Fe), photocatalytic (PAN_TiO2) and their combination (PAN_TiO2_Fe) catalysts.
Nanomaterials 10 00148 g010
Table 1. Process parameters for the membranes production.
Table 1. Process parameters for the membranes production.
MaterialsFlow Rate (mL/h)Voltage (kV)Electrode Distance (cm)Relative Humidity (%)Deposition Time (min)
PAN support215 ÷ 2020 ÷ 2530 ÷ 40120
Catalysts deposition223 ÷ 251520 ÷ 3060
Table 2. Catalyst content of the membranes.
Table 2. Catalyst content of the membranes.
MembraneCatalyst Content (mg/cm2)
PAN_TiO20.14
PAN_Fe0.12
PAN_TiO2_Fe0.19
(53% TiO2 + 47% FeSO4)

Share and Cite

MDPI and ACS Style

Bonora, R.; Boaretti, C.; Campea, L.; Roso, M.; Martucci, A.; Modesti, M.; Lorenzetti, A. Combined AOPs for Formaldehyde Degradation Using Heterogeneous Nanostructured Catalysts. Nanomaterials 2020, 10, 148. https://doi.org/10.3390/nano10010148

AMA Style

Bonora R, Boaretti C, Campea L, Roso M, Martucci A, Modesti M, Lorenzetti A. Combined AOPs for Formaldehyde Degradation Using Heterogeneous Nanostructured Catalysts. Nanomaterials. 2020; 10(1):148. https://doi.org/10.3390/nano10010148

Chicago/Turabian Style

Bonora, Renato, Carlo Boaretti, Laura Campea, Martina Roso, Alessandro Martucci, Michele Modesti, and Alessandra Lorenzetti. 2020. "Combined AOPs for Formaldehyde Degradation Using Heterogeneous Nanostructured Catalysts" Nanomaterials 10, no. 1: 148. https://doi.org/10.3390/nano10010148

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop