Photocatalytic Degradation of H S in The Gas-Phase Using A Continuous Flow Reactor Coated With Tio - Based Acrylic Paint

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 37

Environmental Technology

ISSN: 0959-3330 (Print) 1479-487X (Online) Journal homepage: http://www.tandfonline.com/loi/tent20

Photocatalytic degradation of H2S in the gas-phase


using a continuous flow reactor coated with TiO2-
based acrylic paint

Eduardo Borges Lied, Camilo Freddy Mendoza Morejon, Rodrigo Leonardo


de Oliveira Basso, Ana Paula Trevisan, Paulo Rodrigo Stival Bittencourt &
Fábio Luiz Fronza

To cite this article: Eduardo Borges Lied, Camilo Freddy Mendoza Morejon, Rodrigo Leonardo de
Oliveira Basso, Ana Paula Trevisan, Paulo Rodrigo Stival Bittencourt & Fábio Luiz Fronza (2018):
Photocatalytic degradation of H2S in the gas-phase using a continuous flow reactor coated with
TiO2-based acrylic paint, Environmental Technology, DOI: 10.1080/09593330.2018.1440010

To link to this article: https://doi.org/10.1080/09593330.2018.1440010

Accepted author version posted online: 13


Feb 2018.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tent20
Publisher: Taylor & Francis & Informa UK Limited, trading as Taylor & Francis

Group

Journal: Environmental Technology

DOI: 10.1080/09593330.2018.1440010

Photocatalytic degradation of H2S in the gas-phase using a continuous


flow reactor coated with TiO2-based acrylic paint

Eduardo Borges Lieda,b*, Camilo Freddy Mendoza Morejonb, Rodrigo


Leonardo de Oliveira Bassoc, Ana Paula Trevisand, Paulo Rodrigo Stival
Bittencourte and Fábio Luiz Fronzaa

a
Department of Biological and Environmental Sciences, Federal University of
Technology – Paraná, Medianeira, Brazil; bPostgraduate Program of Chemical
Engineering, West Paraná State University, Toledo, Brazil; cCenter for Natural
Sciences, Federal University for Latin American Integration, Foz do Iguaçu, Brazil;
d
Postgraduate Program of Agricultural Engineering, West Paraná State University,
Cascavel, Brazil; eDepartment of Chemistry, Federal University of Technology –
Paraná, Medianeira, Brazil

*E-mail: lied.eduardo@gmail.com; phone number: +55 (45) 98402 8181; postal adress:
Universidade Tecnológica Federal do Paraná, Câmpus Medianeira, Av. Brasil, 4232.
CEP 85884-000 – Caixa Postal 271 – Medianeira, PR, Brasil

For the photocatalytic degradation of the hydrogen sulphide (H2S) in the gas-
phase it was developed a rectangular reactor, coated with acrylic paint supported
on fiber cement material. The surface formed by the paint coverage was
characterized structural and morphologically by scanning electron microscopy
with energy dispersive X-ray (SEM-EDX) and X-ray diffraction (XRD) analysis.
The flow rate and the inlet concentration of H2S were evaluated as operational
performance parameters of the reactor. Removal efficiencies of up to 94% were
obtained at a flow rate of 2 L min-1 (residence time of 115 s) and inlet
concentration of 31 ppm of H2S. In addition, the H2S degradation kinetics was
modelled according to the Langmuir-Hinshelwood (L-H) model for the inlet
concentrations of 8 to 23 ppm of H2S. The results suggest that flow rate has a
more important influence on photocatalytic degradation than the inlet
concentration. It is assumed that H2S has been oxidized to SO42-, a condition that
led to a deactivation of the photocatalyst after 193 minutes of semi-continuous
use.

Keywords: Photocatalytic paint; heterogeneous photocatalysis; fiber cement;


reaction kinetics; TiO2-rutile

Introduction

The hydrogen sulphide (H2S) is a corrosive and toxic compound that has a

characteristic smell of rotten egg whose perception limit corresponds to a concentration

of approximately 0.5 ppb [1, 2]. H2S can be released as a by-product of various

processes, from animal recycling industries [3,4], to meat processing industries [5], oil

refining, production of cellulose and paper [6], waste treatment plants [7] and others.

Conventional methods commonly used to treat H2S emissions and other

odoriferous compounds. These methods include biofiltration [8-10], chemical cleansing

[11,12] or the adsorption technologies [13-15]. Despite some favorable aspects, these

methods do not have long-term stability. Also, they are non destructive and require

chemicals or high energy consumption [6, 16].

In this regard, Yu et al. [17], Li et al. [18] and Kataoka et al. [2] point out that

the development of an efficient H2S removal technique is highly desired, becoming the

focus of interest in many studies. Furthermore, the growing public concern about

environmental pollution, has served as a vector for the development of unconventional

pollutant removal technologies, such as advanced oxidation processes (AOPs). In the

last years, it is possible to verify great interest in the oxidation processes, especially
through heterogeneous photocatalysis, due to its potential to degrade a wide range of

gaseous pollutants without significant energy consumption. According to Chen and

Poon [19] the main advantage of photocatalysis is the need only of photonic energy to

activate the chemical conversions, a characteristic that contrasts with the conventional

catalysis reactions that require heat for its activation.

In this context, attention should be given to TiO2 because it is the most studied

semiconductor for photocatalytic applications and also because it is the oxide that

presents the best balance of efficiency, chemical stability, low toxicity and price. These

attributes are fundamental requirements for gas-phase photocatalysis applications, since

they avoid the need to install costly filtration processes [6].

Studies about heterogeneous photocatalysis applied to the gas-phase pollutants

show expressive numbers, which reflects the technical feasibility of the use of this

photochemical phenomenon for the treatment of pollutants [2, 16, 20-27].

Additionally, there are many studies about the degradation of hydrogen

sulphide (H2S) using heterogeneous photocatalysis [1, 17, 18, 24-27] in which it is seen

as a promising technique, once the authors of those studies have been able to achieve

degradation efficiencies of up to 100% of the concentration of H2S.

In addition to hydrogen sulphide, photocatalytic degradation with TiO2 can be

used on other gaseous compounds, such as dimethyl sulphide [28-31, 79],

dimethyldisulphide [22, 29, 30], n-octane, n-hexane and n-decane [23], acetaldehyde

and ammonia [26], ethylene [31], phenol [32] and others.

In most of the studies carried out in photocatalysis the TiO2 is used in the form

of powdered particles. On the other hand, for many applications, the most suitable form

is a thin film strongly bonded to a substrate such as glass or ceramics [24, 33-35]. The

use of TiO2 for photocatalysis has a wide and diverse application, and this
semiconductor can be incorporated into formulations of other products, in order to make

its development and application more technically feasible [36, 37].

Conventional paints use TiO2 as the pigment in their formulation. Thus,

depending on certain characteristics the paints may superficially present photocatalytic

properties [36, 38, 39]. There are several reports that show the application of paints with

photocatalytic properties for the treatment of gaseous pollutants, differing mainly with

respect to the type of paint used, and most used types are acrylic paint [38, 40, 41],

vinyl-based paint [42-44] and liquid suspension paint [45].

According to several works, the main substrate used for impregnation of

photocatalysts is ceramic or cement material, although the literature points out different

processes that can be used to connect the photocatalyst to the surface. However, as

already said, most photocatalytic applications for materials involve the mixing of TiO2

into concrete, mortars and ceramic materials [19, 41, 46-48]. As photocatalysis is a

phenomenon that happens on a surface, solutions such as coating (paints) are also

interesting because they can be applied on existing structures and at the same time,

result in lower consumption of TiO2 [41].

The present paper reports an experimental study of photocatalytic degradation of

H2S in the gas-phase using a rectangular reactor coated with white TiO2-based acrylic

paint supported on fiber cement (ceramic). The coating (photocatalyst) on the support

material was characterized in order to obtain information about the surface morphology,

elemental and molecular composition, besides identifying the crystallographic phase of

TiO2 present in the photocatalyst. The performance of the reactor was evaluated respect

to the inlet concentration of H2S and the flow rate. In addition, the kinetics of the

photocatalytic degradation of H2S was adjusted to the Langmuir-Hinshelwood models.


This way the work has relevance in the direction that the study of effective and

economically feasible techniques in the treatment of gaseous compounds has been

highlighted as an important aspect in the area of environmental decontamination.

Experimental part

Experimental apparatus

The experimental module for the photocatalytic degradation assays of H2S is

shown schematically in Figure 1. The arrangement of this experimental apparatus

consists of the following parts: air compressor (1); flow regulator (2); flowmeter (3);

container with H2S generating solution (4); valve (5); photocatalytic reactor (6);

photocatalyst (Fotosan® + fiber cement) (7); transparent surface (8); UV lamp (9); UV

closed chamber (10); silicone hose (11); H2S monitoring system (12); analyser housing

entrance (13); analyser housing exit (14); Alphasense® electrochemical sensor, H2S-BE

model (15); computer (16); wifi communication (17); analyser housing (18).

[FIGURE 1 NEAR HERE]

Figure 1. Schematic representation of the experimental apparatus for the tests of


photocatalytic degradation.

Concentrations of H2S, in the inlet flow, may be variable and are obtained by

altering the pH of the sulphide solution or by the dilution achieved by varying the inlet

gas flow rate. Ultraviolet radiation is emitted by four mercury vapor lamps of 250 W

without the external bulb.

The concentrations of H2S at the inlet and outlet of the piston flow reactor were

measured using an analyzer equipped with an Alphasense® electrochemical sensor

(H2S-BE sensor) and the amount of H2S degraded from the gas-phase was evaluated by

using the Equation 1:


[ ] [ ]
(%) = × 100 (1)
[ ]

Where [ ] and [ ] are the concentrations (ppm) of H2S at the inlet and outlet

of the reactor, respectively.

Photocatalyst

The photocatalyst was obtained from commercial acrylic white paint supported

on a fragment of fiber cement material (Figure 2). The photocatalyst was used without

further pretreatment and all chemicals used were made from analytical grade reagents.

Distilled water was used in the preparation of solutions.

[FIGURE 2 NEAR HERE]

Figure 2. (a) Upper face with acrylic white paint (photocatalyst) supported on fiber
cement material and (b) bottom face without paint coating on a 5x magnification.

The photocatalyst was used to cover 0,0435 m2 of surface area (upper face) on

the support of fiber cement material, as shown in Figure 2 (a). The bottom face (Figure

2 (b)) was not coated with the photocatalyst.

Generation of hydrogen sulphide

The generation of hydrogen sulphide followed the procedures described by

Canela et al. [1]. The experiment for gas generation consists in the pH control of a

solution of sodium sulphide (Na2S.9H2O) and sodium phosphate (Na2HPO4). The

solution was prepared by the addition of 0,6 g of sodium sulphide and 1,44 g of sodium

phosphate in a volume of 100 mL of distilled water. Under these experimental

conditions, when the airflow produced by the compressor passes through the

“headspace” of the solution container, an atmosphere of H2S is generated and conducted


to the photocatalytic reactor.

Modelling the kinetics of H2S degradation

The study of the kinetics of the photocatalytic degradation reaction of H2S was

carried out based on the Langmuir-Hinshelwood (L-H) model, as follows:

=− (2)

Where k is the reaction rate constant; K is the adsorption constant; C is the H2S

concentration; and r is the reaction rate.

In order to do so the experimental data (previously collected) were incorporated

into the model, and the quality of its fit to the linearized model was analyzed (R2),

according to equation 3:

1 1 1 1
= + ∙ (3)

Where Ceq is the H2S concentration in the equilibrium. Thus, according to equation 3

the parts should have good linearity.

Reaction rate per unit of surface area

The reaction rate per unit area can be defined as shown by the Equation 4:

[ ] [ ]
= (4)

Where r is the rate of reaction per unit area; Q is the gas flow rate; [ ] is the H2S

concentration at the reactor inlet; [ ] is the concentration of H2S at the outlet of

the reactor; A is the surface area impregnated with photocatalyst.


Characterization techniques

XRD

The X-ray diffraction (XRD) analysis were performed using a X’Pert Pro MPD

diffractometer from Panalytical with X-ray tube with Cu target (wavelength (λ) = 1,54

Å), 40kV and 45mA. Angular 2θ range from 10° to 120° was used.

The identification of the crystalline phases was done by using the software X-

PertHighScore (Panalytical) and the identified phases were obtained in the Inorganic

Crystal Structure Database (ICSD). The crystallite size was calculated by the well

known Scherrer equation (Equation 5):

= (5)

Where D is the particle size, k is the particle form factor, λ is the wavelength of the

radiation to be used (1,54 Å), β is the full width at half maximum (FWHM) of the peak,

and θ is the diffraction angle.

In order to perform the analysis it was necessary a previous spraying the paint,

subjected to a drying process. After that, it was made the subsequent spraying in an

appropriate granulometry for the XRD tests.

FTIR

Molecular analysis of the surface of the photocatalyst was performed using the

Fourier Transform Infrared Spectroscopy (FTIR) technique. It was necessary to prepare

the painted sample in a form of pellet to be analysed by FTIR. The pellet consists of a

mixture of a transparent matrix in which the sample is attached. The alkaline powder

used was potassium bromide (KBr). The analyses were conducted on an Infrared

Spectrometer, model MB-100 from Bomem.


The assay that determines the materials by infrared spectroscopy was performed

according to ASTM E1252 standards. A small amount of the sample coating was

removed using a sharp blade and then macerated with KBr (optical accessory). The

mixture was pelletized on a hydraulic press at 7000 kg of load. The FTIR measurement

was carried out using the following conditions: acquisition range of 4000 to 400 cm-1; 4

cm-1 resolution; 16 scans.

SEM-EDX

The morphological characteristics of the photocatalyst surface were evaluated by

scanning electron microscopy (SEM). Qualitative chemical analisys was made using an

energy dispersive X-ray (SEM-EDX). The equipment used for the chemical and

morphological characterization of the surface was, respectively, the scanning electron

microscope Tescan, model VEGA 3, with X-ray dispersive energy detector Oxford.

In order to obtain better images at large magnifications. The sample was coated

with an ultra-thin layer of 10 nm of gold to obtain an electrically conductive surface

layer

Results and discussion

Figures 3 (a) and (b) show the X-ray diffractograms and the infrared spectra,

respectively, for the acrylic paint sample.

[FIGURE 3 NEAR HERE]

Figure 3. Results of (a) X-ray diffraction and (b) infrared spectroscopy of the acrylic
paint sample.

The interpretation of the diffractogram of the acrylic paint sample shown in

Figure 3 (a) indicates that the pigment that composes the paint is related to the
crystallographic phase TiO2-rutile with a peak of greater intensity in 2θ = 27°, as also

interpreted by Nogueira [49]. No other crystalline TiO2 phase was identified. The

crystal structure of rutile consists in TiO6 octahedra bonded by the vertices (corner-

shared) in a tetragonal cell [50]. Peaks of lower intensity in 2θ = 36°, 41º, 54º and 69º

also refers to rutile phase. Also, Borges et al. [51] when analyzing a painted sample did

not identify the anatase crystalline phase, only the rutile phase which is more

photostable. Thus, the results of XRD confirm that commercial paints have their

formulation based on the use of TiO2-rutile pigment [37, 52].

Also, according to Figure 3 (a) the low intensity peaks are characteristic of

impurities present in the sample, which may be due to spurious phases detected by X-

ray fluorescence analysis in % of mass, just as like as Al2O3 (0,86), SiO2 (1,03), Cl

(0,06), K2O (0,10), CaO (0,25) and Fe2O3 (0,55).

The size of the TiO2-rutile crystallite in the acrylic paint is approximately 46 nm

(Scherrer equation), with a density of 4,23 g cm-3. Allen et al. [36] calculated the

density and obtained the value of 4,0 g cm-3. In other studies, the TiO2-rutile particle

size was found to be 74 nm [53], 70 nm [33], 42 nm [54] and 40 nm [55]. The

thermodynamically stable phase is related to crystallite sizes greater than 35 nm [34].

In Figure 3 (b) the FTIR spectra of the coating paint of the fiber cement material

fragment are shown. The values and intensities found in the infrared spectrum are

expressed in terms of transmittance (%). The coating paint showed predominant

chemical composition of polyester resin. In fact, the main bands observed are, mainly

due to the resin, in 1730 cm-1 of the ester group (C=O), 1159 cm-1 (C–O), 698 cm-1 (C-

H aromatic substitution) and 2920 cm-1 (stretch-CH aromatic). Confirmation of the

aromatic ring (C = C stretch) is done in the region of 1600-1500 cm-1.


According to Canela et al. [1] there is also another band attributed to the

vibrational enlargement of the OH groups surface bound to the Ti atoms (≡Ti-OH) near

the 3600 cm-1 region. These results confirm the presence of hydroxyl ions in the

structure of the photocatalyst, and this fact is of great relevance considering that these

ions constitute oxidizing agents in the photocatalytic phenomena [56-58, 64, 78].

The gravimetric procedure for the impregnation of acrylic paint on the fiber

cement material yielded an approximate mass amount of 4,7 g. The degree of

incorporation of the photocatalyst (paint) to the support (fiber cement) reached 1,61% in

mass.

The SEM-EDX characterization technique was used to visualize the morphology

and topography, as well as the elemental composition of the surface formed in the fiber

cement material. Electron micrographs and EDX spectra are shown in Figures 4 and 5,

respectively.

[FIGURE 4 NEAR HERE]

Figure 4. Micrograph of the surface of the fiber cement material without the
impregnation of acrylic paint with magnification of (a) 20x and (c) 1650x and
micrograph of the same surface impregnated with acrylic paint using magnification of
(b) 20x and (d) 1650x.

Observation of the microstructure by SEM (Figure 4) shows that the surface has

a large amount of pore sizes (Figure 4 (c)) in the range of 20-800 µM. The acrylic paint

was able to form a film and partially cover the cavities and pores (Figure 4 (d)) in the

fiber cement material used as carrier (Figure 4 (c)).

Generally, in the paints the resin surrounding the pigment particles deforms

during the film forming process, however in some cases the resin cannot deform

sufficiently so as to form a pore-free film and as a result the TiO2-Resin interface


presents regions of higher exposure of TiO2 particles and emergence of TiO2-rutile

particles that was observed by Tryba et al. [59] while also using acrylic paint. The

emergence of pigment particles is a favorable factor for the use of the photocatalytic

properties of the paint.

Considering this possibility of exploitation of the photocatalytic properties

Marolt et al. [38] points out that the development of photocatalytic paints raises at least

two fundamental questions that need to be addressed: (a) Are TiO2 nanoparticles

integrated into the paints sufficiently exposed to irradiation to exhibit measurable

photocatalytic activity? (b) Will the integrated TiO2 nanoparticles destroy the polymeric

(resin) components of the paint?

Therefore, regarding to the matter of TiO2 being sufficiently exposed, it is

possible that the deformation of the previously mentioned film increases the degradation

process of the polymeric components as mentioned by Marolt et al. [38]. According to

the authors, the TiO2 particles are covered with binder (resin), and after irradiation the

TiO2 nanoparticles on the surface are exposed and the photocatalytic activity of the

paint is substantially increased.

Based on this, Marolt et al. [38] clarifies that there is a more efficient and

certainly more reproducible way to increase photocatalytic activity. This method

involves merely changing the content of the organic binders (resin), since the intention

would be to make the initial content of the binder insignificant in the activation of the

photocatalytic process.

In this regard, it should be noted that acrylic paint used to forming the

photocatalytic film shown in Figures 4 (b) and (d) were diluted in water in a ratio of 10:

1, that reduces the resin composition on the volume of paint


In addition, according to the information given by the paint manufacturer, the

TiO2 content in the paint is high, approximately 48% of TiO2, which gives it the visual

appearance of matte paint, a feature which explains why matte paints have a high

pigment volume concentration (PVC). = × 100.

Generally, the concentration of pigment in paint formulations reached values around

18% [42, 45, 52] and 10% [40]. Thus, when the TiO2 concentration in the paint

formulation is high (40 to 50%), the resin fraction becomes insufficient to completely

fill the voids between the TiO2 particles, so the film becomes weaker and porous. Sousa

et al. [52] also reported that the paint used in their study showed a concentration of

pigment volume above the critical value, so the surface became very porous, since high

content of pigment proportionally means low resin content, and as the resin is

responsible for the permeability of the paint, the higher the pigment content the more

porous the film, which according to Ângelo et al. [43] makes the access of the

photocatalyst to the pollutant easier. On contrary, the lower the pigment content, less

porous the film is and more efficient to protection.

The influence of PVC was also investigated by Tryba et al. [59], where the

authors concluded that non-photocatalytic organic paints had a much lower amount of

pigmentary TiO2 than other photocatalytic paints, and therefore non-photocatalytic

paints had a rather high amount of carbon, which in this case would be the composition

of the binder (resin).

Figure 5 (a) shows the EDX spectrum of the surface of the fiber cement

supporter without the paint coating, while Figure 5 (b) shows the spectrum of the same

surface with the paint coating. As expected, Ti is present in the sample with paint. The

detection of Al and Si in greater intensity in Figure 5 (b) only reveals the use of these

substances as extenders (or fillers) in the paint formulation.


[FIGURE 5 NEAR HERE]

Figure 5. EDS spectra of the surface of the fiber cement material (a) without and (b)
with the impregnation of acrylic paint.

The TiO2 (pigment) present in the ink formulation of this work was probably

produced from the sulphonation process [34, 36] as it identifies the sulfur peak (S) in

the spectrum of Figure 5 (B). This process has the intermediary TiOSO4, so the

presence of S can be the result of traces of the sulphate ion on the TiO2 surface. These

traces of sulfur may eventually reduce the photocatalytic properties of TiO2, however

this assumption can only be confirmed by tests of photocatalytic activity, although the

studies by Noda et al. [60] and Noda et al. [61] showed that photocatalytic activity

remains significant.

Similar to the present work, the EDX analysis of photocatalytic paints

performed by Tryba et al. [59] revealed the presence of Ca, Mg, C, O, Ti, Si, K, Al and

trace of S.

Figure 6 shows the mapping of chemical elements present in the sample of

acrylic paint + fiber cement, made by the EDX technique, where the elements are

represented by different colors.

[FIGURE 6 NEAR HERE]

Figure 6. EDX images representing concentration profiles for component elements of a


given region of the photocatalytic surface of the fiber cement with acrylic paint: (a)
electron micrograph; (b) distribution of titanium, calcium, silicon, aluminium and
oxygen in a region of the sample; (c) titanium; and (d) calcium.

The mapping of elements by EDX image aims mainly to show the distribution of

titanium (Ti) and calcium (Ca). It is possible to identify from the images that the

titanium presented higher concentration in the cavity regions of the surface of the fiber
cement (Figures 6 (b) and (c)), so it is inferred that the formation of the paint film was

affected by the morphology of the support, because in cavity regions the presence of Ti

occurs in higher concentration, and in higher topography regions the concentration of

Ca (characteristic feature of the cement support) was predominant (Figure 6 (d)),

although these regions show concentrations of Ti, but in smaller proportions, which

may indicate that in these regions the paint film is thinner.

Results of Photolysis Test, Absence of UV Radiation and Photocatalytic Activity

The results of the photolysis test, the experiment with absence of UV radiation

(in the dark) and the photocatalytic activity assay are presented in Figures 7 (a), (b) and

(c), respectively. The objective was to evaluate the possible existence of H2S

degradation phenomenon only with the presence of UV radiation and the influence of

phenomena that may occur in the absence of UV radiation as well as the characteristic

occurrence of photocatalytic activity, respectively.

[FIGURE 7 NEAR HERE]

Figure 7. Experiments carried out to verify the effects of (a) photolysis for UV
radiation; (b) absence of UV radiation; and (c) photocatalytic activity (Q = 5 L min-1).

Figure 7 (a) shows that direct irradiation test did not lead to the reduction of the

[ ]
H2S concentration values ( [ ] = 1), and the small variations over

time are attributed to a small instability of the H2S generating solution itself and the air

drag system. Therefore, it is concluded that the effect of UV radiation is practically

negligible for hydrogen sulphide degradation.

In other works [1, 24, 62] the authors also report that photolysis has negligible

or even non-existent effect, since they suggest and prove in their studies that H2S has
low UV absorption. Barnes et al. [63] pointed that absorption by H2S occurs in the

region between 190 and 270 nm. The incident radiation supplied by the ultraviolet lamp

used in the present work, has its emission spectrum between 290 and 400 nm, according

to the manufacturer's information.

Before carrying out the photocatalytic activity assays, one must be sure that the

step related to the study of the phenomenon is the exclusive chemical reaction of

photocatalysis and not the adhesion of H2S mass to the catalytic surface. Therefore, it is

necessary to understand the possible effects of adsorption and kinetics of photocatalytic

reaction that can occur simultaneously.

From this purpose, possible evidence of the effect of adsorption was investigated

using low flow rates. The results expressed in Figure 7 (b) suggest that in the tested

range - 2, 5 and 8 L min-1 (RT = 155 s, 46 s and 28 s, respectively) - no significant

effect of lowering H2S concentration was found as a function of time. This was

attributed to some interaction phenomenon (e.g. adsorption) of H2S on the catalytic

surface in the absence of radiation. Thus, the photocatalytic activity assays started

without the need to perform any procedure related to the establishment of H2S

adsorption equilibrium.

As can be seen in Figure 7 (c), the concentrations of H2S gradually decreased

with time after the beginning of "UV on" irradiation, and soon the H2S concentrations

are restored during "UV off". This behavior suggests the significant occurrence of H2S

degradation by the photocatalysis reaction. The photocatalytic phenomenon is explained

by photogeneration of electrons (e-) and gaps (h+) induced by the semiconductor

excitation, so the pair e-/h+ can react with compounds in contact with the surface of the

photocatalyst and initiate oxidation or reduction reactions [50, 58, 64, 66]. Canela et al.
[1] and Kato et al. [16] proposed the complete oxidation of H2S to sulphate through the

following mechanism:

+ ℎ (6)
°
ℎ + → + (7)
°
ℎ + → (8)
+2 +2 → (9)
°
+8 → +2 +4 (10)
+4 → +2 +4 (11)

According to Portela et al. [6] sulphate formation represents an inherent primary

problem, since it accumulates on the surface of the catalyst, covering the active sites

and leading to a progressive deactivation of the catalyst, aspects that will be discussed

in the chapter "Deactivation of the photocatalyst".

Modelling the kinetics of photocatalytic degradation of H2S

In Figure 8 it is shown the L-H model adjustment for H2S degradation using the

acrylic-based photocatalyst supported on fiber cement material.

[FIGURE 8 NEAR HERE]

Figure 8. Adjustment of kinetic experimental data to the Langmuir-Hinshelwood model.

With regard to the L-H model, a good adjustment (R2 ≈ 0.94) of the

experimental results was obtained, with values of 7,7 ppm min-1 and 0,03 ppm-1 for k

and K, respectively. This indicates that L-H expression can be used to describe the rate

of photocatalytic degradation of H2S. The literature reports several papers that use the

L-H model as the main assumption to describe the mechanisms involved in the

photocatalytic degradation [24, 66, 67, 42].


According to this model, the degradation rate of the reagents is only the product

of an apparent rate constant and the Langmuir adsorption term, and therefore the mass

transfer phenomenon does not give any control to the general photodegradation kinetics

[26].

Evaluation of the effect of the volumetric flow rate

The volumetric flow rate was investigated in the range of 2 to 14 L min-1, while

other operating conditions, such as H2S concentration, relative humidity, temperature,

photocatalyst dose and UV radiation were kept constant. H2S removal efficiencies were

calculated using Equation 1 and the reaction rates determined by Equation 4. The results

are shown in Figures 9 (a) and (b).

[FIGURE 9 NEAR HERE]

Figure 9. (a) Conversion of H2S as a function of the inlet flow rate to the reactor; (b)
reaction rate per unit area as a function of the inlet flow.

Figures 9 (a) and (b) show the results of photocatalytic oxidation using different

flow rates, where is evident that lower flow rate results in higher conversion. The

maximum removal efficiency of 94% was obtained with the flow rate of 2 L min-1

(Figure 9 (a)), which corresponds to a theoretical residence time of approximately 1 min

and 55 s. The results of Figure 9 (a) indicate that, increasing the flow rate to 5 L min-1,

the conversion can be greater than 85%. This means that mass transfer limitations do

not affect very strongly under these conditions. However, from flow rates above 5 L

min-1 the conversion decreased significantly as the flow increased, reaching a

conversion of 33% to a flow rate of 14 L min-1.

Some authors [22, 24, 25, 27, 41] have reported the same behavior and attributed

the reduced conversion to a problem of contact time of the contaminant with the
reaction surface, a condition that is verified in different ranges of flow rates, as is the

case of Guillard et al. [22] ~ 0,005 to 0,09 L min-1, Brancher et al. [24] 0,4 to 2,5 L min-
1
, Martinez et al. [41] 1 to 5 L min-1, Alonso-Tellez et al. [27] 0,1 to 0,98 L min-1, Kako

et al. [25] 0,096 to 1,25 L min-1, Jo [68] 1 to 4 L min-1 and Yu & Brouwers [69] 1 to 5 L

min-1. In summary, these authors conclude that lower flow rates result in greater

efficiencies in H2S conversion, a favorable condition due to the longer contact time.

Even under higher operating conditions, the residence time is reduced, so this time

would be insufficient for the contact between the gas stream and the photocatalyst,

reducing the efficiency of the process.

It is possible to understand these results based on the literature [70, 71, 66].

These authors have reported that the effect of the flow rate on the photocatalytic

degradation of compounds in the gas-phase can be explained in one of the hypotheses

set forth in the discussion.

In a first situation, in low flow rate ranges, the increase of the flow contributes to

the increase of the efficiency of the process, because in theory, increasing the flow rate

would increase the rate of convective mass transfer [72]. This demonstrates that mass

transfer to the catalyst surface may limit the process. In this sense, the theoretical

number of Reynolds = was calculated for flow rates from 2

to 14 L min-1 ( = 3,7.10-3 a 2,6.10-2 m s-1), corresponding to the Reynolds

numbers from 24 to 174, so that the flow in the reactor can be described as laminar (Re

<2100). This suggests that reaction rates may theoretically be dependent on mass

transfer, although the observed results point to a negative correlation between the H2S

degradation efficiency and the flow rate. This indicates that the reaction rate was

controlled by the kinetics, regardless of the the Reynolds number, conclusions also

reported with even smaller values of Reynolds number, such as: 2 to 19 [71], 29 to 59
[72], 21 to 105 [70].The hypothesis that mass transfer are not relevant to the process is

strengthened by the analyses of the fluid dynamics of the reactor obtained by computer

simulation (CFD) using the software Comsol Multiphysics 5.0. The results, is the speed

profiles (m s-1) that can be seen in Figures 10 (a) and (b).

[FIGURE 10 NEAR HERE]

Figure 10. Example of internal velocity profile (m s-1) to the photocatalytic reactor in
different cutting planes: (a) three-dimensional view “xyz”, and (b) side view “zy” (vinlet
= 0,528 m s-1; Q = 5 L min-1).

From Figures 10 (a) and (b) it is evident that flow preferably takes place in the

lower and central portions of the reactor (occurrence of preferential flow). Therefore,

based on the simulation results the Reynolds numbers ( = )

were recalculated considering the speed rates obtained in the simulation

( çã =0,12 to 0,73 m s-1). The obtained values of correspond to the

range of 152 to 928, which are above to the previously calculated values of

. Obviously, the precise description of the fundamental principles of mass

transfer and reaction kinetics occurring simultaneously in the reactors is a complex

issue. So, the use of literature references on the subject has its relevance in the sense

that the statements and conclusions of the literature serve as parameters for the results

of this work.

In this sense, it is possible to demonstrate, based on other authors, that the mass

transfer phenomena did not produce effects on the reaction speed on the studied

conditions of this work. Zhang et al. [73] were able to demonstrate that reaction rate

constants were smaller than the mass transfer coefficients. For these results, speeds

were above 0,60 m s-1, so the authors concluded that mass transfer effects were not
important under those conditions. In addition, Obee & Brown [74] also reported that the

effects of mass transfer were not significant at speeds above 0,10 m s-1.

In a second hypothesis, in which the values of flow rates are intermediate, its

variation does not have considerable theoretical effect in the efficiency, which means

that the phenomenon is controlled by the kinetics of superficial reaction. However, the

results of Figures 9 (a) and (b) show that the behaviour is likely to more accurately

express a third hypothesis, which occurs at ranges that exhibit relatively high flow rates.

Therefore, the increased flow rate decreases the residence time for the transfer of

contaminants from the gas-phase to the surface of the catalyst, which implies in

decreasing efficiency.

Evaluation of the effect of the inlet concentration

The effect of the inlet concentration was investigated in the range of 7 to 23

ppm, while other operating conditions such as flow, relative humidity, temperature and

UV radiation were kept constant. The H2S degradation efficiencies were calculated

using Equation 1 and the reaction rates determined by Equation 8. The respective

results are shown in Figures 11 (a) and (b).

[FIGURE 11 NEAR HERE]

Figure 11. (a) Conversion of H2S as a function of the i concentration variation on the
reactor; and (b) reaction rate per unit area as a function of affluent concentration.

Figures 11 (a) and (b) express the effect of the inlet concentration on the H2S

degradation rate using acrylic paint. The degradation rate (degradation efficiency) vs

inlet concentration, as denoted in Figure 11 (b), approaches a first-order reaction, so the

conversion tends to remain constant with the change in inlet concentration (8 to 23


ppm), as shown in Figure 11 (a), although the inlet concentration has slightly affected

the studied values.

This result is consistent with other authors [1, 18, 23], both for low and high

values of inlet concentration. Using TiO2 fixed on support of nickel material, Li et al.

[18] studied the influence of the inlet concentration on the H2S degradation in a range

from 53 to 408 ppb. The conversion values obtained for the studied range showed that

inlet concentration affected the conversion efficiency in a low intensity manner.

Concerning the fact that the conversion was little affected, other authors report

that the concentration did not affect the conversion, as is the case of Rochetto and

Tomaz [23], who studied the influence of the concentration on the n-octane degradation,

using values between 60 and 110 ppm in the incoming current, with a residence time of

24 s. The conversion values obtained for the studied range showed that the inlet

concentration did not affect the conversion efficiency.

In another study Canela et al. [1] also showed no decrease in reaction rate with

increasing H2S concentration even at high concentrations. The authors evaluated the

oxidation rate for H2S as a function of the input concentration (30 to 855 ppm) and a

flow rate of 0,2 L min-1. The reaction was found to be of a first-order, such that

photocatalytic degradation of H2S in the concentration range was maintained constant at

approximately 99%, regardless of feed concentration.

Deactivation and proposed reaction mechanism for H2S degradation

In order to evaluate the time of deactivation of the photocatalyst, the time

intervals in which the photocatalyst was used in the experiments without significant loss

of efficiency were quantified, for this it was considered that deactivation was reached in

a condition that efficiency was below 10%. Thus, it was possible to determine that

deactivation of the photocatalyst occurred after 193 min (3,21 hours) of semi-
continuous use.

The deactivation of the photocatalyst by the photocatalytic oxidation of H2S is

reported in other works [1, 27] where the authors characterized the formation of

photocatalytic reaction product accumulation on the photocatalyst by the infrared

spectroscopy technique. In other similar works [16, 25] the authors were able to obtain

characterization results by X-ray excitation photoelectron spectroscopy after the

degradation tests.

In general, as already mentioned in the text, the literature proposes that the

mechanism for photocatalytic degradation is that H2S sulfur atom was mainly oxidized

to sulphate ion (SO42-) and trapped in the photocatalyst: + +2 →

( ) + [16].

The accumulation of SO42- on the surface of the photocatalyst indicates that this

factor is the reason to the deactivation of TiO2, constituting a problem inherent for the

photocatalytic degradation of H2S. Kataoka et al. [2] reinforces this hypothesis by

explaining that the SO42- ion does not exist in the gaseous phase and presumably these

compounds would accumulate on the surface.

Yu et al. [17] reported that 66% of the sulfur atoms of the destroyed H2S were

recovered as sulphate. The authors did not identify peaks of elemental sulfur (S),

hydrogen sulphide (H2S) and sulfur dioxide (SO2). The characterization results used by

these authors suggest that the main product of the photocatalytic degradation of H2S is

SO4-2.

Conclusions

Photolysis experiments showed that only UV radiation was not able to promote

H2S degradation. Photocatalytic degradation efficiencies of 94% were obtained for a


flow rate of 2 L min-1 (115 s) and inlet H2S concentration of 31 ppm. For the flow rate

of 2 L min-1, the velocity per unit of calculated surface area was 0,042 mol m-2.min-1,

while for 14 L min-1 the velocity reached a value of 0,017 mol m-2 min-1. The results

suggest that the flow has more influence on the degradation than the inlet concentration.

The kinetic study showed good adjustments for the L-H model, so it was

considered that these models adequately represent the kinetics of H2S photocatalytic

degradation. The constants of the L-H model were estimated in k = 7,7 ppm min-1 and K

= 0,03 ppm-1.

From the realization of the deactivation process it is assumed that the H2S was

oxidized to SO42-. As the literature suggests, the SO42- species accumulate gradually on

the surface during the photocatalytic reaction causing the photocatalyst to undergo a

deactivation process throughout its use, which in this work took place in 193 min.

References

[1] Canela MC, Alberici RM, Jardim WF. Gas-phase destruction of H2S using
TiO2/UV-VIS. Journal of Photochemistry and Photobiology A: Chemistry. 1998;
112:73-80.

[2] Kataoka S, Lee E, Tejedor-Tejedor MI, Anderson MA. Photocatalytic degradation


of hydrogen sulfide and in situ FT-IR analysis of reaction products on surface of TiO2.
Applied Catalysis B: Environmental. 2005; 61:159-163.

[3] Defoer N, Bo ID, Langenhove HV, Dewulf J, Elst TV. Gas chromatography–mass
spectrometry as a tool for estimating odour concentrations of biofilter effluents at
aerobic composting and rendering plants. Journal of Chromatography A. 2002; 970:
259–273.

[4] Anet B, Lemasle M, Couriol C, Lendormi T, Amrane A, Cloirec PL, Cogny G,


Fillières R. Characterization of gaseous odorous emissions from a rendering plant by
GC/MS and treatment by biofiltration. Journal of Environmental Management. 2013;
128: 981–987.

[5] Rappert S, Muller R. Odor compounds in waste gas emissions from agricultural
operations and food industries. Waste Management. 2005; 25: 887–907.
[6] Portela R, Suárez S, Rasmussen SB, Arconada N, Castro Y, Durán A, Ávila P,
Coronado JM, Sánchez B. Photocatalytic-based strategies for H2S elimination. Catalysis
Today. 2010; 151:64-70.

[7] Tsang YF, Chua H, Sin SN, Chan SY. Treatment of odorous volatile fatty acids
using a biotrickling filter. Bioresource Technology. 2008; 99: 589–595.

[8] Ramírez M, Fernández M, Granada C, Borgne SL, Gómez JM, Cantero D.


Biofiltration of reduced sulphur compounds and community analysis of sulphur-
oxidizing bacteria. Bioresource Technology. 2011; 102: 4047–4053.

[9] Luo J, Lindsey S. The use of pine bark and natural zeolite as biofilter media to
remove animal rendering process odours. Bioresource Technology. 2006; 97: 1461–
1469.

[10] Sheridan BA, Curran TP, V.A. Dodd. Biofiltration of n-butyric acid for the control
of odour. Bioresource Technology. 2003; 89: 199–205.

[11] Ma Y, Chen Z, Gong H. Study on selective hydrogen sulfide removal over carbon
dioxide by catalytic oxidative absorption method with chelated iron as the catalyst.
Renewable Energy. 2016; 96: 1119–1126.

[12] Pope D, Davis BJ, Moss RL. Multi-stage absorption of rendering plant odours
using sodium hypochlorite and other reagents. Atmospheric Environment. 1980; 15:
251–262.

[13] Grandclerc A, Guéguen-Minerbe M, Nour I, Dangla P, Chaussadent T. Impact of


cement composition on the adsorption of hydrogen sulphide and its subsequent
oxidation onto cementitious material surfaces. Construction and Building Materials.
2017; 152: 576–586.

[14] Mochizuki T, Kubota M, Matsuda H, Camacho LFD. Adsorption behaviors of


ammonia and hydrogen sulfide on activated carbon prepared from petroleum coke by
KOH chemical activation. Fuel Processing Technology. 2016; 144: 164–169.

[15] Boraphech P, Thiravetyan P. Trimethylamine (fishy odor) adsorption by


biomaterials: Effect of fatty acids, alkanes, and aromatic compounds in waxes. Journal
of Hazardous Materials. 2015; 284: 269–277.

[16] Kato S, Hirano Y, Iwata M, Sano T, Takeuchi K, Matsuzawa S. Photocatalytic


degradation of gaseous sulfur compounds by silver-deposited titanium dioxide. Applied
Catalysis B: Environmental. 2005; 57:109-115.

[17] Yu Y, Zhang T, Zheng L, Yu J. Photocatalytic degradation of hydrogen sulfide


using TiO2 film under microwave electrodeless discharge lamp irradiation. Chemical
Engineering Journal. 2013; 225:9-15.

[18] Li X, Zhang G, Pan H. Experimental study on ozone photolytic and photocatalytic


degradation of H2S using continuous flow mode. Journal of Hazardous Materials. 2012;
199–200: 255–261.
[19] Chen J, Poon C. Photocatalytic construction and building materials: From
fundamentals to applications. Building and Environment. 2009; 44:1899-1906.

[20] Boyjoo Y, Sun H, Liu J, Pareek VK, Wang S. A review on photocatalysis for air
treatment: From catalyst development to reactor design. Chemical Engineering Journal.
2017; 310: 537-559.

[21] Fu X, Zeltner WA, Anderson MA. Applications in photocatalytic purification of


air. Studies in Surface Science and Catalysis. 1996; 103: 445-461.

[22] Guillard C, Baldassare D, Duchamp C, Ghazzal MN, Daniele S. Photocatalytic


degradation and mineralization of a malodorous compound (dimethyldisulfide) using a
continuous flow reactor. Catalysis Today. 2007; 122:160-167.

[23] Rochetto UL, Tomaz E. Degradation of volatile organic compounds in the gas
phase by heterogeneous photocatalysis with titanium dioxide/ultraviolet light. Journal of
the Air & Waste Management Association. 2015; 65:810-817.

[24] Brancher M, Franco D, Lisboa HM. Photocatalytic oxidation of H2S in the gas
phase over TiO2-coated glass fiber filter. Environmental Technology. 2016; 37: 2852-
2864.

[25] Kako T, Nakajima A, Watanabe T, Hashimoto K. Comparison of photocatalytic


properties of a batch reactor with those of a flow reactor in a nearly controlled mass
transport region. Research on Chemical Intermediates. 2005; 31:371-378.

[26] Sopyan I. Kinetic analysis on photocatalytic degradation of gaseous acetaldehyde,


ammonia and hydrogen sulfide on nanosized porous TiO2 films. Science and
Technology of Advanced Materials. 2007; 8:33-39.

[27] Alonso-Tellez A, Robert D, Keller N, Keller V. A parametric study of the UV-A


photocatalytic oxidation of H2S over TiO2. Applied Catalysis B: Environmental. 2012;
115-116: 209-21.

[28] Langenhove HV, Demeestere K, Dewulf J, Witte B. Titanium dioxide mediated


heterogeneous photocatalytic degradation of gaseous dimethyl sulfide: Parameter study
and reaction pathways. Applied Catalysis B: Environmental. 2005; 60:93-106.

[29] Cantau C, Larribau S, Pigot T, Simon M, Maurette MT, Lacombe S. Oxidation of


nauseous sulfur compounds by photocatalysis or photosensitization. Catalysis Today.
2007; 122:27-38.

[30] Nishikawa H, Takahara Y. Adsorption and photocatalytic decomposition of odor


compounds containing sulfur using TiO2/SiO2 bead. Journal of Molecular Catalysis A:
Chemical. 2001; 172:247-251.

[31] Einaga H, Tokura J, Teraoka Y, Ito K. Kinetic analysis of TiO2-catalyzed


heterogeneous photocatalytic oxidation of ethylene using computational fluid dynamics.
Chemical Engineering Journal. 2015; 263:325-335.
[32] Turki A, Guillard C, Dappozze F, Ksibi Z, Berhault G, Kochkar H. Phenol
photocatalytic degradation over anisotropic TiO2 nanomaterials: Kinetic study,
adsorption isotherms and formal mechanisms. Applied Catalysis B: Environmental.
2015; 163:404-414.

[33] Feltrin J, Sartor MN, De Noni Jr A, Bernardin AM, Hotza D, Labrincha JA.
Photocatalytic surfaces of titania on ceramic substrates. Part II: substrates, deposition
and heat treating processes. Cerâmica. 2014; 60:1-9.

[34] Feltrin J, Sartor MN, De Noni Jr A, Bernardin AM, Hotza D, Labrincha JA.
Photocatalytic surfaces of titania on ceramic substrates. Part I: Synthesis, structure and
photoactivity. Cerâmica. 2013; 59:620-632.

[35] Saleiro GT, Cardoso SL, Toledo R, Holanda JNF. Evaluation of crystalline phases
of titanium dioxide supported on red ceramics. Cerâmica. 2010; 56:162-167.

[36] Allen NS, Edge M, Verran J, Stratton J, Maltby J, Bygott C. Photocatalytic titania
based surfaces: Environmental benefits. Polymer Degradation and Stability. 2008;
93:1632–1646.

[37] Hochmannova L, Vytrasova J. Photocatalytic and antimicrobial effects of interior


paints. Progress in Organic Coatings. 2010; 67:1-5.

[38] Marolt T, Škapin AS, Bernard J, Živec P, Gaberšček M. Photocatalytic activity of


anatase-containing facade coatings. Surface & Coatings Technology. 2011; 206:1355–
1361.

[39] Auvinen J, Wirtanen L. The influence of photocatalytic interior paints on indoor air
quality. Atmospheric Environment. 2008; 42:4101-4112.

[40] Maggos T, Bartzis JG, Liakou M, Gobin, C. Photocatalytic degradation of NOx


gases using TiO2-containing paint: A real scale study. Journal of Hazardous Materials.
2007; 146:668-673.

[41] Martinez T, Bertron A, Ringot E, Escadeillas G. Degradation of NO using


photocatalytic coatings applied to different substrates. Building and Environment. 2011;
46:1808-1816.

[42] Monteiro RAR, Lopes FVS, Silva AMT, Ângelo J, Silva GV, Mendes AM,
Boaventura RAR, Vilar VJP. Are TiO2-based exterior paints useful catalysts for gas-
phase photooxidation processes? A case study on n-decane abatement for air
detoxification. Applied Catalysis B: Environmental. 2014; 147:988-999.

[43] Ângelo J, Andrade L, Mendes A. Highly active photocatalytic paint for NOx
abatement under real-outdoor conditions. Applied Catalysis A: General. 2014; 484:17–
25.
[44] Águia C, Ângelo J, Madeira LM, Mendes A. Influence of photocatalytic paint
components on the photoactivity of P25 towards NO abatement. Catalysis Today. 2010;
151: 77–83.

[45] Costa A, Chiarello GL, Selli E, Guarino M. Effects of TiO2 based photocatalytic
paint on concentrations and emissions of pollutants and on animal performance in a
swine weaning unit. Journal of Environmental Management. 2012; 96:86-90.

[46] Ballari MM, Hunger M, Hüsken G, Brouwers HJH. NOx photocatalytic


degradation employing concrete pavement containing titanium dioxide. Applied
Catalysis B: Environmental. 2010; 95:245-254.

[47] Wu Y, Krishnan P, Yu LE, Zhang M. Using lightweight cement composite and


photocatalytic coating to reduce cooling energy consumption of buildings. Construction
and Building Materials. 2017; 145:555-564.

[48] Silva AL, Muche DNF, Dey S, Hotza D, Castro RHR. Photocatalytic Nb2O5-doped
TiO2 nanoparticles for glazed ceramic tiles. Ceramics International. 2016; 42:5113-
5122.

[49] Nogueira MV. Photocatalysts based on titanium dioxide modified with niobium to
reduce carbon dioxide to methanol [master´s thesis]. Araraquara: Paulista State
University; 2014.

[50] Lan Y, Lu Y, Ren Z. Mini review on photocatalysis of titanium dioxide


nanoparticles and their solar applications. Nano Energy. 2013; 2:1031-1045.

[51] Borges SS, Xavier LPS, Silva AC, Aquino SF. Immobilized titanium dioxide
(TiO2) in different support materials to use in heterogeneous photocatalysis. Química
Nova. 2016; 39: 836-844.

[52] Sousa VM, Manaia CM, Mendes A, Nunes OC. Photoinactivation of various
antibiotic resistant strains of Escherichia coli using a paint coat. Journal of
Photochemistry and Photobiology A: Chemistry. 2013; 251:148-153.

[53] Wan G, Wang S, Li L, Mu G, Yin X, Zhang X, Tang Y, Yi L. Photocarrier


dynamic measurement of rutile TiO2 films prepared by RF magnetron reactive
sputtering. Journal of Alloys and Compounds. 2017; 701:549-553.

[54] Ziarati A, Badiei A, Ziarani GM, Eskandarloo H. Simultaneous photocatalytic and


catalytic activity of p–n junction NiO@anatase/rutile-TiO2 as a noble-metal free
reusable nanoparticle for synthesis of organic compounds. Catalysis Communications.
2017; 95:77-82.

[55] Allen NS, Edge M, Ortega A, Sandoval G, Liauw CM, Verran J, Stratton J,
McIntyre RB. Degradation and stabilisation of polymers and coatings: nano versus
pigmentary titania particles. Polymer Degradation and Stability. 2004; 85:927-946.

[56] Fujishima A, Zhang X. Titanium dioxide photocatalysis: present situation and


future approaches. Comptes Rendus Chimie. 2006; 9:750-760.
[57] Fujishima A, Rao TN, Tryk DA. Titanium dioxide photocatalysis. Journal of
Photochemistry and Photobiology C: Photochemistry Reviews. 2000; 1:1–21.

[58] Ochiai T, Fujishima A. Photoelectrochemical properties of TiO2 photocatalyst and


its applications for environmental purification. Journal of Photochemistry and
Photobiology C: Photochemistry Reviews. 2012; 13:247-262.

[59] Tryba B, Homa P, Wróbel RJ, Morawski AW. Photocatalytic decomposition of


benzo-[a]-pyrene on the surface of acrylic, latex and mineral paints. Influence of paint
composition. Journal of Photochemistry and Photobiology A: Chemistry.2014; 286:10–
15.

[60] Noda LK, Almeida RM, Probst LFD, Gonçalves NS. Characterization of sulfated
TiO2 prepared by the sol–gel method and its catalytic activity in the n-hexane
isomerization reaction. Journal of Molecular Catalysis A: Chemical. 2005; 225:39-46.

[61] Noda LK, Almeida RM, Gonçalves NS, Probst LFD, Sala O. TiO2 with a high
sulfate content—thermogravimetric analysis, determination of acid sites by infrared
spectroscopy and catalytic activity. Catalysis Today. 2003; 85:69-74.

[62] Huang L, Xia L, Ge X, Jing H, Dong W, Hou H. Removal of H2S from gas stream
using combined plasma photolysis technique at atmospheric pressure. Chemosphere.
2012; 88: 229–234

[63] Barnes AJ, Hallam HE, Howells JDR. Photolysis of hydrogen sulphide in low-
temperature matrices. J Mol Struct. 1974; 23:463–467.

[64] Fujishima A, Zhang X, Tryk DA. TiO2 photocatalysis and related surface
phenomena. Surface Science Reports. 2008; 63: 515–582.

[65] Nakata K, Fujishima A. TiO2 photocatalysis: Design and applications. Journal of


Photochemistry and Photobiology C: Photochemistry Reviews. 2012; 13: 169–189.

[66] Doucet N, Bocquillon F, Zahraa O, Bouchy M. Kinetics of photocatalytic VOCs


abatement in a standardized reactor. Chemosphere. 2006; 65: 1188–1196.

[67] Asenjo NG, Santamaría R, Blanco C, Granda M, Patricia Álvarez, Rosa Menéndez.
Correct use of the Langmuir–Hinshelwood equation for proving the absence of a
synergy effect in the photocatalytic degradation of phenol on a suspended mixture of
titania and activated carbon. Carbon. 2013; 55: 62–69.

[68] Jo W. Coupling of titania with multiwall carbon nanotubes for decomposition of


gas-phase pollutants under simulated indoor conditions. Journal of the Air & Waste
Management Association. 2013; 63: 963–970.

[69] Yu QL, Brouwers HJH. Indoor air purification using heterogeneous photocatalytic
oxidation. Part I: Experimental study. Applied Catalysis B: Environmental. 2009; 92:
454–461.
[70] Shayegan Z, Lee C-S, Haghighat F. TiO2 photocatalyst for removal of volatile
organic compounds in gas phase – A review. Chemical Engineering Journal. 2017; doi:
https://doi.org/10.1016/j.cej.2017.09.153

[71] Alberici RM, Jardim WF. Photocatalytic destruction of VOCs in the gas-phase
using titanium dioxide. Applied Catalysis B: Environmental. 1997; 14: 55–68.

[72] Verbruggen SW, Lenaerts S, Denys S. Analytic versus CFD approach for kinetic
modeling of gas phase photocatalysis. Chemical Engineering Journal. 2015; 262: 1–8.

[73] Zhang Y, Yang R, Zhao R. A model for analyzing the performance of


photocatalytic air cleaner in removing volatile organic compounds. Atmospheric
Environment. 2003; 37: 3395–3399.

[74] Obee TN, Brown RT. TiO2 photocatalysis for indoor air applications: effects of
humidity and trace concentration levels on the oxidation rates of formaldehyde, toluene
and 1,3-butadiene. Environmental Science and Technology. 1995; 29: 1223–1231.

You might also like