Você está na página 1de 10

Journal of Photochemistry and Photobiology A: Chemistry 306 (2015) 2130

Contents lists available at ScienceDirect

Journal of Photochemistry and Photobiology A:


Chemistry
journal homepage: www.elsevier.com/locate/jphotochem

Simultaneous removal of two industrial dyes by adsorption and


photocatalysis on a y-ashTiO2 composite
Anca Duta, Maria Visa *
Transilvania University of Brasov, R&D Center, Renewable Energy Systems and Recycling, Eroilor 29, 500036 Brasov, Romania

A R T I C L E I N F O

A B S T R A C T

Article history:
Received 13 January 2015
Received in revised form 25 February 2015
Accepted 4 March 2015
Available online 14 March 2015

A novel composite was obtained in mild hydrothermal conditions using y ash, TiO2 and a cationic
surfactant (HTAB). The components were involved in extensive re-structuring processes, evidenced by
XRD, FT-IR, SEM, AFM, BET and surface energy measurements. The composite was used as adsorption
substrate and as photocatalyst for the simultaneous removal of two commercial dyes (Bemacid Red, BR;
Bemacid Blue, BB). The experimental tests were run at the natural pH of the suspension, which was
alkaline (pH 10.6). In the experimental conditions electrostatic interactions due are less extensive, and
the data can be directly linked to the substrates morphology and dyes molecular structure and size. It
was found that adsorption efciency is strongly inuenced by the micro-pores on the substrate and the
kinetics mainly depends on the dyes exibility. For the azo dye (BR), photocatalysis was found to run
independent from adsorption while for the anthraquinone dye (BB) the process proves to be more
complex. Under UV irradiation, the removal efciencies reached after 240 min were of 93% for BR and of
77% for BB.
2015 Elsevier B.V. All rights reserved.

Keywords:
Fly ash TiO2 composite
Adsorption
Photodegradation
Dyes mixture

1. Introduction
Pollution was identied as one of the major threats to
humanity; therefore more and more investigations are dedicated
to implementing the concept of removing waste by using wastes.
Water is one of the most affected environmental factor and the
waste by waste concept is now largely applied for wastewater
treatment using solid wastes resulted from usual goods as LCD
devices [26] or from highly polluting industries as alumina
processing, [13] or steel metallurgy [14]. Energy production based
on solid, carbon-based fuels was long ago recognized to pose huge
environmental problems not only as result of the gaseous products
but also because of the solid by-product, the ash. Thus special
attention was devoted to identifying various effective paths to use
the ash: as construction material [18], mainly by partially replacing
sand in concrete [37]. Despite all the effort, only about 40% of the
ash is used for cement production, therefore alternative solutions
are looked after. The ash results as coarse grains removed from the
bottom of the furnace (bottom ash) and ne particles captured by
the electrolters from the gaseous exhaust (y ash, FA). Both
fraction were tested as adsorbents for pollutants removal from
wastewaters; bottom ash was reported effective for dyes removal

* Corresponding author. Tel.: +40 729109355.


E-mail address: maria.visa@unitbv.ro (M. Visa).
http://dx.doi.org/10.1016/j.jphotochem.2015.03.007
1010-6030/ 2015 Elsevier B.V. All rights reserved.

[22] but more interest is raised by y ash, as this has a much larger
specic surface. As such, y ash is a moderate adsorbent and has
the disadvantage of variable composition (inuenced by the type
of coal, burning process and burning equipment). Therefore many
alternatives are reported to make y ash a more reproducible
substrate and to increase its surface area and charge [1]. Fly ash
with optimized surface properties was reported as an efcient
adsorbent for anionic [39], basic [28] or organo-reactive dyes [11].
A step forward was done when using y ash in hydrothermal
synthesis for obtaining zeolites or zeolite-type structures, which
proved to be highly efcient in dyes adsorption [46,23,33]. The
addition of various inorganic compounds allowed tailoring the
adsorbent properties [48], while the addition of cationic surfactants led to substrates with improved adsorption efciency [27].
As advanced oxidation processes, especially heterogeneous
photocatalysis emerged as promising alternatives to the current
wastewater treatment to remove recalcitrant pollutants, the use of
y ash was considered mainly as substrate for the photocatalyst
deposition. Recent reports are on the synthesis of the photocatalyst
from a precursor system directly on nano-sheets [17] and on the
y-ash grains/cenospheres [47,45]. These materials are using FA as
an inert substrate and the organics degradation is expected only
due to the photocatalytic effect. Lately, the active adsorption effect
of natural clay was reported combined with photocatalysis [41],
and previous work outlined that the y ashTiO2 mix can
effectively activate in adsorptionphotocatalysis [44].

22

A. Duta, M. Visa / Journal of Photochemistry and Photobiology A: Chemistry 306 (2015) 2130

For dyes removal from wastewaters, most of the papers report


the results obtained from mono-pollutant solutions; but, industrial
wastewaters contain more than one dye that can interact in the
aqueous environment and can develop concurrent processes on
the active sites, therefore this type of data is required for an
accurate process design.
This paper reports on the simultaneous adsorption and
photodegradation of two industrial dyes with signicantly
different molecular structure and stability to degradation (Bemacid Blue and Bemacid Red), using a novel composite material based
on y ash and TiO2 obtained in mild hydrothermal synthesis.
A novel approach of the possible concurrent and parallel processes
during dyes removal on the zeolite-type structure is presented,
based on the comparative adsorption and photocatalysis kinetic
data; the importance of the dye molecular structure and of the
substrates surface aspect is outlined, along with the possible
oxidationreduction reactions, involving the aqueous species and
the photocatalytic substrate.

agent, able to regulate zeolite-type structures. The cationic


surfactant was added in a concentration lower than CMC measured
in the reaction system (CMCHTAB = 298 mg/L, evaluated based on
conductivity measurements). This value is very close to that
experimentally measured for HTAB in water (CMCHTAB = 294 mg/L)
and shows that in the alkaline environment there is little
adsorption of the surfactant on the solid grains in the slurry, thus
direct contamination is limited.
The FATiO2 composite was obtained under stirring
(300 rot/min) for 48 h, in a reaction vessel with reux condenser,
at 100  C and atmospheric pressure. After the reactions were
completed, the suspended matter was washed with ultra-pure
water till constant pH, ltered, dried at 105115  C and further
treated at 235  C for 24 h to remove the residual organic
compounds. The sample was denoted FAA-DCS.
The novel nano-composite material was tested as adsorption
substrate and photocatalyst in the simultaneous treatment of
synthetic wastewaters dyes-loaded.

2. Experimental

2.2. Characterization of the novel composite material

Fly ash was collected from the electro-lters of the CHP plant in
Brasov, Romania. The main oxides in the y ash composition (as
given by the CHP plant) are: SiO2, Al2O3 and Fe2O3; according to the
ASTM standards (C618-05), the y ash is of F type (pozzolanic),
thus during long contact with water there is no aggregation. The
SiO2:Al2O3 ratio (2:41) is quite large and shows that FA represents a
possible zeolite precursor. Several other oxides of interest were
identied in the y ash composition, Table 1, and are important as
possible participants in photodegradation as heterogeneous
photocatalysts (TiO2) or in the development of in situ photoFenton systems (Fe2O3 and Mn2O3); the y ash also contains
unburned carbon, representing possible adsorption sites for dyes.

The FAA-DCS crystalline structure was evaluated by XRD


(Advanced D8 Discover Bruker diffractometer, CuKa1 = 1.5406 ,
40 kW, 20 mA, 2u range 10 . . . 70 , scanning step 0.02 , scan
speed 2 s/step) and complementary data were obtained by FTIR
spectroscopy (Spectrum BX PerkinElmer BX II 75548, l = 400
4000 nm).
The surface characterization included micro-porosity and BET
specic surface measurements (Autosorb-IQ-MP, Quantachrome
Instruments), morphology studies before and after dyes removal
(roughness and macro-pores size distribution) using AFM (Ntegra
Spectra, NT-MDT model BL222RNTE, in semi-contact mode with
Golden silicon cantilever, NCSG10, at constant force 0.15 N/m, with
a 10 nm tip radius) and scanning electron microscopy (SEM, S3400N-Hitachi, accelerating voltage of 20 KV). Surface elemental
composition was evaluated using energy dispersive X-ray spectroscopy (EDS, Thermo Scientic Ultra Dry) using the sessile drop
method (OCA-20 Contact Angle-meter, Data Physics Instruments).

2.1. Materials and nano-composite preparation


As the data in Table 1 shows, y ash (FA) contains also soluble
oxides therefore, their removal by washing represents a compulsory step to get a substrate with constant properties; therefore, raw
y ash was washed with ultra-pure water (1:1 mass ratio), under
mechanical stirring (100 rpm, Nahita GJ-1 stirrer), at room
temperature (22  1  C), for 48 h, followed by ltering, drying at
105115  C for 4 h, and sieving to collect the 2040 mm fraction
(FAw, representing 37% from the total ash). This substrate is highly
heterogeneous and has limited adsorption efciency, proved in
preliminary tests (up to 20%).
Following the concept for developing a novel sustainable and
lower costs substrate, active in adsorption and photocatalysis, FAw
was mixed with TiO2 (Degussa P25, Evonik), a nanosized powder,
containing 71% anatase, 27% rutile and 2% amorphous [34]. The
powders were mixed in a previously optimised ratio FAw:
TiO2 = 3:1, [42]. The cationic surfactant hexadecyltrimethylammonium bromide (HTAB) was added to the slurry prepared in NaOH
2N [43], for surface charge control and as potential templating

Table 1
Fly ash composition.
Major oxides (%)
SiO2
53.32

Al2O3
22.05

Fe2O3
8.97

CaO
5.24

MgO
2.44

K2O
2.66

Na2O
0.63

TiO2
1.07

MnO
0.08

LOI
1.58

V
115

Mn
800

Co
12

Trace elements (ppm)


Ba
700

Cu
60

Zr
100

Sn
3

LOI: unburned carbon.

Pb
35

As
100

Ni
55

Zn
160

Cr
100

Ti
>3000

2.3. Adsorption and photocatalytic experiments


The chemical structures of the dyes are presented in Fig. 1. The
chromophore groups are of azo type in BR and of anthraquinone
type in BB, both resistant to biodegradation.
The pollutant systems were synthetically prepared using
bidistilled water; the pollutants were added as the dyes Bemacid
Blau (BB) and Bemacid Rot (BR), (Bezema AG, Switzerland), used as
received.
Experiments were done on systems containing a dyes mix of BB
50 mg/L + BR 50 mg/L. Two types of processes were investigated:
- Adsorption on the nano-composite substrate (FAA-DCS): batch

adsorption experiments were done under visible light, at room


temperature (22  1  C), by mixing 0.5 g substrate with 50 mL
solution under mechanical stirring. The adsorption mechanism
and the kinetic data were evaluated; comparative experiments
run in dark showed that the substrates are not Vis-active as
photocatalysts. These experiments are denoted with (A).
- Photodegradation investigations were done on FAA-DCS suspensions with the same composition as in the adsorption
studies, with and without hydrogen peroxide (30%) addition,
under UV irradiation, and the results are denoted with (F). The
H2O2 concentration in the system was 0.6 g/L. The homemade
photocatalytic reactor is equipped with three F18W/T8 black
tubes (Philips), emitting UV-A light in the region of 340400 nm
and lmax(emission) = 365 nm. The mean value of the radiation

A. Duta, M. Visa / Journal of Photochemistry and Photobiology A: Chemistry 306 (2015) 2130

23

Fig. 1. Chemical structure of the dyes.

ux intensity, reaching the middle of the reacting suspension,


measured with a digital Luxmeter (Mavolux5032C/BUSM) was
3Lx. During the experiments quartz beakers were used.
All the experiments were run at the natural pH of the system
(pH 10.6) when no precipitation or turbidity was observed.
During the kinetic studies, aliquots were taken at xed
moments (up to 240 min) when stirring was briey interrupted
and, after ltration on 0.45 mm lter, the supernatant was
analysed. Preliminary experiments proved that dyes losses due
to adsorption on the beaker walls or on the ltering paper were
negligible.
The dyes concentration before and after removal was evaluated
based on UVvis absorbance spectra (PerkinElmer UVvis spectrophotometer, Lambda-950); in the dyes mixtures, overlapping is
observed along with and a shift of the correspondent single dyes
peaks, therefore the concentrations were calculated after calibration, based on the rst order derivative of the absorbance spectrum
of each dye in the mix, as previously described [2], at the maximum
absorption wavelength of BB (lBB = 652 nm) and BR (lBR = 444 nm).
The adsorption/photodegradation efciency of the dyes was
evaluated using Eq. (1):

c 0  ct
 100%
c0

(1)

where c0 represents the initial concentration and ct the pollutants


concentration at moment t.
3. Results and discussions
3.1. The nano-composite material
The crystalline structures of washed y ash (FAw) and of the FA
TiO2 nanocomposite were comparatively investigated. The XRD
spectra, Fig. 2, show overall crystalline percentages of 42% in FAw
and 64% in FAA-DCS; the results outline that most of the pristine
and washed y ash (FAw) components are contributing to the
crystalline matrix.
As expected, the composite FAA-DCS contains large amounts of
crystalline TiO2 polymorphs from Degussa P25 (anatase and rutile),
but also preserves the brookite already identied in FAw. Beside
these, the composite also shows new crystalline alumino-silicates
(Al2SiO5, Na6Al6Si10O3212H2O), proving that the long treatment in
mild hydrothermal conditions, led to internal restructuring, mainly
involving the crystalline quartz (as its corresponding peak in
FAA-DCS is much smaller compared to FAw) and crystalline or
amorphous aluminates and/or alumino-silicates. Lowering the
quartz content represents a positive aspect when targeting the

410.43
954.21

2500

Intensity [a.u.]

2000

"

# 0
#

# Na6Al6Si10O32
* Al2O3-SiO2
Anatase syn
0 Quartz syn
o Rutile
" Kyanite Al2SiO5

"

#
0

(2)

1500

1000

Cchaoitehexagonal

FAA-DCS
FAw

TiO2brookite

Fe2O3

0.8

595.21

0.7
0.6
Absorbance [a.u.]

FAA-DCS
0
FAA-DCS t=235 C
FAA-DCS (BB+BR)

3463.52

1639.78

(1)

0.5
1443.74

0.4

(2)

1639.78

0.3

(3)

0.2
0.1

500

(1)

0.0
500

0
10

20

30

40

50

60

70

1000

1500

2000

2500

3000

3500

4000

-1

[cm ]

2thete [degree]

Fig. 2. XRD graphs of (1) FAw; (2) FAA-DCS.

Fig. 3. IR spectra of (1) FAA-DCS before thermal treatment, (2) FAA-DCS, (3) FAADCS after (BB + BR) adsorption.

24

A. Duta, M. Visa / Journal of Photochemistry and Photobiology A: Chemistry 306 (2015) 2130

Table 2
Surface characteristics of FAw and FAA-DCS.
Sample

Specic surface area (BET)


(m2/g)

Average micro-pores
diameter
(nm)

Average macro-pores diameter


(nm)

Average roughness
(nm)

Surface energy
(mN/m)
Polar

Dispersive

FAw
FAA-DCS

6.14
62.18

27.2
13.6

570 . . . 780
280 . . . 360

90.8
135.6

46.09
83.74

23.87
1.27

development of zeolite-type structures as this, along with mullite,


were found to be mainly responsible for hindering the zeolite
formation and activity [23]. Another interesting nding is the
possible dissolution of Fe2O3 during the composite synthesis, thus
the interference of the Fe3+/Fe2+ couple in the surface oxidation
reduction reactions (ORR) is unlikely.
The crystallite sizes of the TiO2 polymorphs were calculated
using the Scherrer formula, Eq. (2):
Kl
(2)
bcosu
where t is the crystallites size; K is the shape factor with a
typical value 0.94; l is the X-ray wavelength (1.5406 ); b is the
line broadening at half of the maximum intensity (of a peak) and u

is the diffraction angle.The results show crystallite sizes of 48 nm


for anatase and of 74 nm for rutile, much larger than the initial

crystallite sizes of Degussa P25 less than 30 nm, [34], and even
larger than the grain sizes (50 nm), proving that the composite is
the result of an internal restructuring of both components
(FAw and TiO2).
To further investigate the composite, FT-IR spectra were
employed, Fig. 3.
Several peaks could be identied, giving evidence for restructuring that mainly affects the tetrahedral external reorganization
and can be linked to an ordering effect leading to zeolite-type
structures; these correspond to [25]: the double ring vibration
(595 cm1), symmetric stretch (764 cm1) and asymmetric stretch
of the SiO or AlO bonds (954 cm1). The peak with less
intensity, recorded at 16351644 cm1, corresponds to the bending
mode in the water molecules; the absorption band observed
at 32003600 cm1 was attributed the hydroxyl group stretching/
vibration in the SiOH, AlOH  Al and SiOH  Al units [32].

Fig. 4. AFM topography of (a) FAw and (b) FAA-DCS.

Fig. 5. SEM images of (a) FAw; (b) FAA-DCS.

A. Duta, M. Visa / Journal of Photochemistry and Photobiology A: Chemistry 306 (2015) 2130

25

Table 3
Elemental surface composition of FAA-DCS.
Element

Na

Atomic % on the bed


Atomic % on the
cenosphere

7.8
9.15 16.9 13.1
8.83 4.92 39.1
0.2
6.69 6.69 11.8
5.37 14.44 6.91 48.07 0.4

Al

Si

Ti

Br

The results also show that post-treating the FAA-DCS composite


did not signicantly alter the silica, alumina and alumina-silicate
bonds.
Obtaining zeolites from y ash was extensively reported during
the past years, [5,1] but, turning it into a possible photocatalyst is
mainly considered by linking/adsorbing TiO2 on cenospheres,
followed by senzitation [35,47]. Our results prove that embedding
TiO2 in the composite combines the advantages of increasing the
crystal ordering of y ash and extends the crystallinity of the most
photo-catalytic active TiO2 polymorph.
As both, adsorption and photocatalysis are surface-dependent
processes, the surface charge and morphology represent important
properties, outlined in Table 2.
The results in Table 2 were obtained by BET experiments
(specic surface and average micro-pores diameter), AFM (roughness and average macro-pores diameter) and surface energy
measurements, using the model developed by Owens, Wendt,
Robel and Kaelble [50].
The washing process removes the soluble oxides leaving a
substrate with low specic surface, rather low roughness and a
morphology with very large open pores. The composite has a large
specic surface (larger than Degussa P25, 50.3 nm), with small
micro-pores (large enough to accommodate the dyes). The reorganizing process is conrmed by the AFM pictures, outlining
signicant differences between the randomly structured surface of
FAw, Fig. 4a, and the regular aggregates with droplet shape, Fig. 4b,
assembled in rough structures that leave open macro-pores on the
FAA-DCS surface.
The predominant polar/ionic surface energy corresponds, in an
oxide material in alkaline pH (larger than the point of zero charge),
to a negatively charged surface and shows that the mild
hydrothermal process increased the surface polarity/ionic degree.
This combination of increased specic surface, homogeneity
and roughness, and negative charges supports the use of the FAADCS composite as substrate in adsorption processes of neutral or
cationic species.
Additional surface investigations were done and the SEM
images are presented in Fig. 5.
The SEM images conrm that the y ash grains were cracked
and micro-restructuring occurred with signicant modication of
the surface aspect, as results of dissolution/re-precipitation
reactions.
The average atomic percentage was evaluated based on EDS.
Although EDS is not a highly accurate overall elemental analysis, it
may give indicative results on the surface composition.
The results in Table 3 outlines, a quite large amount of surface
Na, which is highly unlikely to be in the corresponding soluble
oxide (Na2O), thus being the counter-ion in alumina-silicates,
including the zeolite-type structures.
The EDS analysis shows different compositions on the cenospheres, with majority composition of alumino-silicates and
carbon; the results also show that titania (partially) coated the
substrate. On the beds, one interesting result is the rather low
oxygen content; as no other titanium based compounds are
expected but TiO2 (consuming about 26.2 units from the total
39.1 oxygen units), rough calculations show that the remaining Si:
Al:O ratio could be estimated at 16.9:9.15:12.8. This oxygen decit
indicates inorganic polymeric structures (as kyanite outlined by

Fig. 6. Adsorption efciency of BR and BB from their mixture, on (b) FAA-DCS and
(b) TiO2.

XRD), but even so, the content is too low. Therefore one might
consider that nitrogen (from HTAB) could also be embedded in
polymeric surface structures, as it also may act as dopant for TiO2.
Considering this oxygen decit it is also highly possible that carbon
is in its elemental form, which represents an advantage in the
further adsorption processes.
3.2. Adsorption processes on FAA-DCS
The adsorption efciency of the dyes from their mixture on
FAA-DCS is quite different, Fig. 6a, proving concurrent processes:
The kinetic parameters were best tted by the pseudo-second
order kinetic model [19], with the linear form given by Eq. (3):
t
1
t

qt k2 q2e qe

(3)

where k2 is the pseudo second-order rate constant (g mg1 min1).

Table 4
Kinetic parameters of the BB, BR dyes removal in adsorption processes.
Pollutant

k2
(g mg1 min1)

qe
(mg g1)

R2

BB (BB + BR) (A)


BR (BR + BB) (A)

35.245
1.408

1.194
4.003

0.961
0.993

26

A. Duta, M. Visa / Journal of Photochemistry and Photobiology A: Chemistry 306 (2015) 2130

Fig. 7. FAA-DCS surface after BB + BR adsorption: (a) AFM image; (b) macro-pores distribution and (c) SEM micrograph.

Based on Eq. (3) the kinetic parameters were calculated and are
presented in Table 4:
The efciency and the kinetic results show that the dye structure,
molecular exibility and dimension differently inuence the
competition for the adsorption sites (on the surface edges and
corners and in the micro-pores). Although large, the BB molecule has
the aromatic rings linked only by s exible bonds, indicating a fast
adsorption process supported by a fast diffusion at the easily
accessiblesites.ThemorerigidBRmolecule(duetotheazo-bond)has
a slower diffusion, thus slower adsorption but able to give use to a
largeramountof adsorptionsites(e.g., inside the small micro-pores).
Comparative investigations were done, in similar conditions, on
the TiO2 substrate, Fig. 6b. The results prove that TiO2 does not
represent a suitable adsorption substrate, as for contact times
longer than 30 min. the predominant process is desorption. The
lower bonding possibilities of BR proves that the adsorption on
FAA-DCS runs on active sites of alumino-silicate or carbon types
and that the contribution of TiO2 in its adsorption is less important.
In these conditions, the TiO2 surface actually remains available to
BB, allowing equilibration at efciencies of about 50% reached after
60 min of contact time.
Additionally, at the working alkaline pH (10.6) electrostatic
interactions are less likely as the substrate is negatively charged
and the anionic dyes cannot involve the SO3 groups. Considering this aspect, the most important property during adsorption is
the amount of active sites and their surface distribution.

The surface aspect after 4 h of adsorption is slightly changed


and conrms a possible preferential adsorption on the large
micropores, as the micro-pores distribution of FAA-DCS show
before and after adsorption, Fig. 7a and b. Also the SEM images,
Fig. 7c, show the adsorption on the surface edges and corners, as
the surface contains a larger amount of small aggregates
(individual or in clusters).
3.3. Photocatalytic processes on FAA-DCS
The active species generated by photo-irradiation will attack
the pollutant, if this is in the very close vicinity of the substrate.
Thus, adsorbed pollutant species will support an efcient photocatalytic processes, and studies are already reported linking the
pollutants structures, adsorption and the photocatalytic efciency
[7,21].
The most active photocatalytic component of the composite
is TiO2. The TiO2 polymorphs have band gaps of 3.0 eV
(rutile), 3.2 eV (anatase) and brookite (with the lowest calculated
direct band gap of 3.54 eV, [49], thus are active under
UV radiation with wavelengths lower than 413 nm for rutile,
387 nm for anatase and, respectively 357 nm for brookite. As the
UV wavelength used in our experiments was 365 nm, we
may conclude that only anatase and rutile are actually activated
under irradiation and can exhibit the well-known coupling
effect [24].

A. Duta, M. Visa / Journal of Photochemistry and Photobiology A: Chemistry 306 (2015) 2130

27

Fig. 8. Photodegradation efciency of BB and BR from dyes mixtures (a) with and (b) without H2O2 addition and (c) comparison of the adsorption (A) and
adsorption + photocatalytic (A + F) processes efciency on FAA-DCS (d) comparison of the adsorption (A) and adsorption + photocatalytic (A + F) processes efciency on TiO2,
after 240 min.

Once formed the electronhole pair, the most common


mechanisms involves the holes for hydroxyl radical production.
In alkaline media several other reactions are possible [12,4,38],
involving the O2/HO, O2/HOO and the O2/H2O2 couples.
Another common and undesired process is the electronhole
recombination, and the photocatalytic efciency strongly depends
on the systems ability to limit this process. Involving the
photogenerated electrons in different reactions represents one
path which can be designed considering the electrochemical
potential and the hydrogen peroxide addition is expected to
support electron trapping. In the working conditions, the HO
production from the direct homolitic break of the peroxy bond is
unlikely, as higher energy is needed (l < 254 nm).
The substrate material can also interact with the electrons,
contributing to the electron trapping mechanism, by:
- Surface reduction of the Ti4+, Eq. (4):

Ti4+ + 1e ! Ti3+ (4)

with a reduction potential for the conduction band of


E0 = 0.52 V, thus explaining the hydrogen-peroxide anion formation, [29]:
- Lattice defects can contribute to electrons trapping, Eqs. (5)(6),

[7]:
Vo + e ! Vo (5)

Vo + O2 ! Vo + O2  (6)


As the EDS data showed, the FAA-DCS composite shows oxygen
decit and these non-stoichiometry (due to oxygen vacancies) can
increase the superoxide production.
- Involving the dopant impurities states, located between the

conduction and valence band of TiO2 [3]. In the case of the


FAA-DCS the obvious impurities can be nitrogen (from HTAB).
- Involving the other oxides from the FA matrix that can form
tandem or diode type interfaces with TiO2 [10].

28

A. Duta, M. Visa / Journal of Photochemistry and Photobiology A: Chemistry 306 (2015) 2130

Fig. 9. Kinetic data modelling using the pseudo-rst order kinetics for (a) BR in adsorption + photocatalysis; (b) BB in adsorption + photocatalysis; (c) BR in assumed
photocatalysis; (d) BB in assumed photocatalysis.

These two paths will be explored and tailored in the future, as


so far VIS-activity, expected from suitable semiconductor coupling
or doping, was not registered in the photocatalytic experiments
involving FAA-DCS.
Once produced, the HO radicals will attack the organic dye and
several mechanisms are possible, [36]:
1. Hydrogen abstraction: RH + HO ! R + H2O (7)
2. Redox reactions: RH + HO ! RH+ + HO (8)
3. Addition to the p-bonds in the pollutant molecule (9)

In a typical evolution, the organic radicals generated in these


reactions will form peroxyl radicals (ROO) in reaction with oxygen
and will initiate chain reactions that can end up in mineralization
products (CO2 and H2O) or in other stable by-products, sometimes
very difcult to degrade.
In terms of efciency, the results of the experiments run under
UV irradiation are presented in Fig. 8, and show an increased in the
bleaching efciency under UV irradiation, as compared to vis for
both dyes.
Comparative investigations using only TiO2 (Fig. 8d) conrm
that H2O2 addition does not improve the process and that the BR
dye is less well degraded on titania. This outlines again the
importance of adsorption as preliminary step in photocatalysis. As

in the case of adsorption (Fig. 6b), the more exible BB is better


removed on TiO2 (with about 15% higher overall efciencies).
However, considering that both dyes should be simultaneously
removed, the FAA-DCS substrate can be considered more suitable,
particularly in systems without H2O2 and this represents an asset,
both in terms of possible photo-corrosion and considering the
process costs.
On the FAA-DCS substrate the effect of UV irradiation is
signicantly different for the two dyes, supporting the assumption
of very different degradation mechanisms, for the azo-(BR) and
anthraquinone-(BB) dyes:

p-bonds in the azo groups represent the primary


target for the HO radicals (ending up into N2), while the sulfonic
groups will also react to further produce phenols or arenes [31],
and the amino-groups will eventually form ammonia [8].
- In BB, the most likely HO attack is to the NH- and sulfonic
groups leading to a broad range of intermediates derived from
the anthraquinone and mono-arene nuclei [15].
- In BR, the

However, most of the papers report on the azo- and


anthraquinone dyes degradation in acidic media [6] but little
information can be found about the mechanisms and on the
efciency in alkaline environment.

A. Duta, M. Visa / Journal of Photochemistry and Photobiology A: Chemistry 306 (2015) 2130

Another interesting nding is the H2O2 effect in the working


conditions, leading after 240 min to an efciency decrease, thus
supporting the assumption of a fast decomposition in the rst
minutes, followed by adsorption of the by-products that could deactivate part of the photocatalytic sites. A similar effect was
previously reported for the pesticides degradation and was
explained by the chlorinated by-products formed during the
photocatalytic reaction on TiO2 [30]. This explanation could be
valid for BR, also containing chlorine atoms linked to a
mononuclear arene, but cannot support the results observed on
BB. On the other hand, the addition of H2O2 is reported to enhance
the electron-acceptor feature of the dye anthraquinone molecules
[40] but as the experimental results show, this effect was not
observed. Therefore, further investigations were employed on the
processes kinetics.
As the dyes concentrations were relatively low, the kinetic data
were initially modeled using a rst order kinetics derived from the
LangmuirHinshelwood model [16], Eq. (10):
 
C
ln
kobs t
C0

(10)

where C and C0 are the current (at moment t) and initial dye
concentrations, and kobs is the overall observed pseudo rst-order
degradation rate constant.
The results are presented in Fig. 9a and b, and show that the
overall mechanism does not follow the pseudo-rst order model,
largely reported as suitable for describing heterogeneous photocatalysis and assuming that during the photo-process the
adsorption-desorption on the photocatalyst is not disturbed by
the decomposition reactions [9]. One obvious reason is the
composite substrate which might simultaneously support (a) only
adsorption and (b) photocatalysis having as preliminary step
adsorption (in the LangmuirHinshelwood model, the oxidant
species and the pollutants are considered to be both adsorbed prior
to oxidation).
One possible alternative is to have parallel adsorption and
photocatalysis mechanisms, independently running. This assumption was investigated and the adsorption data were subtracted
from the overall photocatalytic results, than, the pseudo-rst order
kinetic model was again applied, Fig. 9c and d. As the graphs show,
the assumption of parallel independent kinetic mechanisms for
adsorption and photocatalysis is well justied for BR. In the
absence of H2O2, the photocatalytic decomposition contribution to
the total removal efciency is of about 14% (Fig. 8c), proving that
the photocatalytic sites are signicantly less than the adsorption
ones. The process runs faster when H2O2 is employed but in these
conditions the overall efciency is similar to adsorption; this
nding is in good agreement with other authors that outlined the
signicant effect of H2O2 on dyes degradation in extreme pH
conditions (pH < 2.5 or pH > 12) but the much lesser effect in the
pH range of 2.5, . . . ,12 [20].
The same approach was developed for the photocatalytic
bleaching of BB from the dyes mixture and the results show that
adsorption and UV activated processes are not running as
independent mechanisms (Fig. 9d). The UV activated processes
are responsible for a much higher share of the total efciency, with
contributions of 57% without H2O2 addition, respectively of 31%
when H2O2 is also part of the system (Fig. 8c). The experimental
results hint that concurrent processes may run, involving the
photodegradation interim products that can occupy/clog or
activate the photocatalytic sites. The BB adsorption proved also
to be faster (k2 = 35.245 g mg1 min1) as compared to BR
(k2 = 1.408 g mg1 min1), thus allowing a more signicant amount
of oxidizing species to be involved. In the pH working conditions,
the oxygen production from H2O2 is employed, which afterwards

29

can contribute to decreasing re-combination. Detailed investigation on specic reaction paths will be developed in the future.
4. Conclusions
wastewaters
are
usually
loaded
with
Industrial
mixed pollutants which can be involved in concurrent or
parallel processes when removed via adsorption or photocatalysis
processes. This is particularly true for dyes, as most of the
shades and colors result from dyes mixes. Therefore, this
paper investigates a mix of two commercial dyes, Bemacid Red
(azo-dye) and Bemacid Blue (anthraquinone-dye), and their
removal by photocatalysis and adsorption. Considering
both processes, a novel composite substrate was developed in
mild hydrothermal conditions starting from y ash and TiO2.
The substrate analyses show that the composite has a regular
surface aspect, high roughness and a signicant specic surface
and charge.
The results show that adsorption efciency signicantly
depends on the surface properties while the process kinetics is
mainly inuenced by the molecular structure (exibility) of the
dye.
The photocatalytic experiments proved various paths for
dyes removal: (a) independent adsorption and photocatalysis
for the azo-dye and (b) a complex mechanism, possibly involving
the adsorption of the intermediate products for the anthraquinone
dye. The addition of H2O2 in the alkaline working conditions
decreased the process efciency and is not recommended. On
the other hand, working in extreme pH conditions to make use
of the hydrogen peroxide would raise sustainability and
cost issues.
These type of kinetic data are needed in the design of the
wastewater treatment processes (and equipment), particularly
when advanced mineralization is targeted, for water re-use.
Acknowledgements
This research was supported by a grant of the Romanian
National Authority for Scientic Research, ANCS UEFISCDI,
project PN-II-PT-PCCA-2013-4-0726. The structural funds project
PRO-DD (ID 123, SMIS 2637, 11/2009) is acknowledged for
providing the infrastructure used in this work.
References
[1] M. Ahmaruzzaman, A review on the utilization of y ash, Prog. Energy
Combust. Sci. 36 (3) (2010) 327363.
[2] L. Andronic, A. Duta, Photodegradation processes in two-dyes systemssimultaneous analysis by rst-order spectra derivative method, Chem. Eng. J.
198199 (2012) 468475.
[3] S. Banerjee, S.C. Pillai, P. Falaras, K.E. Oshea, J.A. Byrne, D.D. Dionysiou, New
insights into the mechanism of visible light photocatalysis, J. Phys. Chem. Lett.
5 (2014) 25432554.
[4] R. Beranek, (Photo) electrochemical methods for the determination of the
band edge positions of TiO2-based nanomaterials, Adv. Phys. Chem. (2011) 20
Article ID 786759.
[5] R.S. Blissett, N.A. Rowson, A review of the multi-component utilisation of coal
y ash, Fuel 97 (2012) 123.
[6] S.H. Chang, K.S. Wang, S.J. Chao, T.H. Peng, L.C. Huang, Degradation of azo and
anthraquinone dyes by a low-cost Fe 0/air process, J. Hazard. Mater. 166 (23)
(2009) 11271133.
[7] L.G. Devi, S.G. Kumar, Exploring the critical dependence of adsorption of
various dyes on the degradation rate using Ln3+TiO2 surface under UV/solar
light, Appl. Surf. Sci. 261 (2012) 137146.
[8] D. El-Mekkawi, H.R. Galal, Removal of a synthetic dye Direct Fast Blue B2RL via
adsorption and photocatalytic degradation using low cost rutile and Degussa
P25 titanium dioxide, J. Hydro-environ. Res. 7 (3) (2013) 219226.
[9] A.V. Emeline, V.N. Kuznetsov, V.K. Ryabchuk, N. Serpone, Heterogeneous
photocatalysis: basic approaches and terminology, in: S.L. Suib (Ed.), New and
Future Developments in Catalysis, Elsevier, Amsterdam, 2013, pp. 147
Chapter 1.

30

A. Duta, M. Visa / Journal of Photochemistry and Photobiology A: Chemistry 306 (2015) 2130

[10] A. Enesca, L. Isac, A. Duta, Charge carriers injection in tandem semiconductors


for dyes mineralization, Appl. Catal. B 162 (2015) 352363.
[11] Z. Eren, F.N. Acar, Adsorption of Reactive Black 5 from an aqueous solution:
equilibrium and kinetic studies, Desalination 194 (13) (2006) 110.
[12] A. Fujishima, T.N. Rao, D.A. Tryk, Titanium dioxide photocatalysis, J.
Photochem. Photobiol. C 1 (1) (2000) 121.
[13] A. Gil, S. Albeniz, S.A. Korili, Valorization of the saline slags generated during
secondary aluminum melting processes as adsorbents for the removal of
heavy metal ions from aqueous solutions, Chem. Eng. J. 251 (2014)
4350.
[14] E.R. Goetz, R.G. Rieer, Performance of steel slag leach beds in acid mine
drainage treatment, Chem. Eng. J. 240 (2014) 579588.
[15] B. Gzmen, B. Kayan, A.M. Gizir, A. Hesenov, Oxidative degradations of reactive
blue 4 dye by different advanced oxidation methods, J. Hazard. Mater. 168 (1)
(2009) 129136.
[16] N. Guetta, H.A. Amar, Photocatalytic oxidation of methyl orange in presence of
titanium dioxide in aqueous suspension. Part II: kinetics study, Desalination
185 (2005) 439448.
[17] P. Hareesh, K.B. Babitha, S. Shukla, Processing y ash stabilized hydrogen
titanate nano-sheets for industrial dye-removal application, J. Hazard. Mater.
229230 (2012) 177182.
[18] M.J.R. Hinojosa, A.P. Galvn, F. Agrela, M. Perianes, A. Barbudo, Potential use of
biomass bottom ash as alternative construction material: conictive chemical
parameters according to technical regulations, Fuel 128 (2014) 248259.
[19] Y.S. Ho, G.J. McKay, Comparative sorption kinetic studies of dye and aromatic
compounds onto y ash, J. Environ. Sci. Health 34 (1999) 11791204.
[20] J. Hong, N. Yuan, Y. Wang, S. Qi, Efcient degradation of Rhodamine B in
microwave-H2O2 system at alkaline pH, Chem. Eng. J. 191 (2012) 364368.
[21] W. Irawatya, F.E. Soetaredjoa, A. Ayucitraa, Understanding the relationship
between organic structure and mineralization rate of TiO2-mediated
photocatalysis, Procedia Chem. 9 (2014) 131138.
[22] C. Jarusiripot, Removal of reactive dye by adsorption over chemical
pretreatment coal based bottom ash, Procedia Chem. 9 (2014) 121130.
[23] B. Jha, D.N. Singh, A three step process for purication of y ash zeolites by
hydrothermal treatment, Appl. Clay Sci. 90 (2014) 122129.
[24] L. Jing, S. Li, S. Song, L. Xue, H. Fu, Investigation on the electron transfer
between anatase and rutile in nano-sized TiO2 by means of surface
photovoltage technique and its effects on the photocatalytic activity, Sol.
Energy Mater. Sol. Cells 92 (9) (2008) 10301036.
[25] O. Keka, C.P. NaraYan, S. Amar Nath, Zeolits from y ash: synthesis and
characterization, Bull. Mater. Sci. 27 (2004) 555564.
[26] C.T. Lee, Preparation of spherical foamed body with function of media for
waste water treatment by using waste LCD glass, J. Ind. Eng. Chem. 20 (5)
(2014) 30893095.
[27] C. Li, Y. Dong, D. Wu, L. Peng, H. Kong, Surfactant modied zeolite as adsorbent
for removal of humic acid from water, Appl. Clay Sci. 52 (4) (2011) 353357.
[28] J.X. Lin, S.L. Zhan, M.H. Fang, X.Q. Qian, H. Yang, Adsorption of basic dye from
aqueous solution onto y ash, J. Environ. Manage. 87 (2008) 193200.
[29] H. Liu, H.T. Ma, X.Z. Li, W.Z. Li, M. Wu, X.H. Bao, The enhancement of TiO2
photocatalytic activity by hydrogen thermal treatment, Chemosphere 50 (1)
(2003) 3946.
[30] S.Y. Lu, D. Wu, O.L. Wang, J. Yan, A.G. Buekens, K.F. Cen, Photocatalytic
decomposition on nano-TiO2: destruction of chloroaromatic compounds,
Chemosphere 82 (9) (2011) 12151224.

[31] M.A. Meetani, S.M. Hisaindee, F. Abdullah, S.S. Ashraf, M.A. Rauf, Liquid
chromatography tandem mass spectrometry analysis of photodegradation of a
diazo compound: a mechanistic study, Chemosphere 80 (4) (2010) 422427.
[32] H. Misran, R. Singh, S. Beguma, M.A. Yarmo, Processing of mesoporous silica
materials (MCM-41) from coal y ash, J. Mater. Process. Technol. 186 (2007)
813.
[33] N.M. Musyoka, L.F. Petrik, O.O. Fatoba, E. Hums, Synthesis of zeolites from coal
y ash using mine waters, Miner. Eng. 53 (2013) 915.
[34] B. Ohtani, O.O. Prieto-Mahaney, D. Li, R. Abe, What is Degussa (Evonik) P25?
Crystalline composition analysis, reconstruction from isolated pure particles
and photocatalytic activity test, J. Photochem. Photobiol. A 216 (23) (2010)
179182.
[35] A.N. kte, D. Karamanis, D. Tuncel, Dual functionality of TiO2yash
nanocomposites: water vapor adsorption and photocatalysis, Catal. Today
230 (2014) 205213.
[36] M.A. Oturan, An ecologically effective water treatment technique using
electrochemically generated hydroxyl radicals for in situ destruction of
organic pollutants: application to herbicide 2,4-D, J. Appl. Electrochem. 30
(2000) 475482.
[37] M. Singh, R. Siddique, Effect of coal bottom ash as partial replacement of sand
on properties of concrete, Resour. Conserv. Recycl. 72 (2013) 2032.
[38] C. Song, J. Zhang, Electrocatalytic oxygen reduction reaction, in: J. Zhang (Ed.),
PEM Fuel Cells Electrocatalysts and Catalyst Layers. Fundamental and
Applications, Springer, 2008, pp. 89133.
[39] D. Sun, X. Zhang, Y. Wu, X. Liu, Adsorption of anionic dyes from aqueous
solution on y ash, J. Hazard. Mater. 181 (2010) 335342.
[40] M.A. Tariq, M. Faisal, M. Saquib, M. Muneer, Heterogeneous photocatalytic
degradation of an anthraquinone and a triphenylmethane dye derivative in
aqueous suspensions of semiconductor, Dyes Pigm. 76 (2) (2008) 358365.
[41] V. Vimonses, B. Jin, C.W.K. Chow, C. Saint, An adsorptionphotocatalysis hybrid
process using multi-functional-nanoporous materials for wastewater
reclamation, Water Res. 44 (2010) 53855397.
[42] M. Visa, L. Andronic, D. Lucaci, A. Duta, Concurrent dyes adsorption and photodegradation on y ash based substrates, Adsorption 17 (2011) 101108.
[43] M. Via, A.-M. Chelaru, Hydrothermally modied y ash for heavy metals and
dyes removal in advanced wastewater treatment, J. Appl. Surf. Sci. (2014) 22.
[44] M. Visa, R.A. Carcel, L. Andronic, A. Duta, Advanced treatment of wastewater
with methyl orange and heavy metals on TiO2, y ash and their mixtures, Catal.
Today 144 (12) (2009) 137142.
[45] B. Wang, C. Li, J. Pang, X. Qing, J. Zhai, Q. Li, Novel polypyrrole-sensitized hollow
TiO2/y ash cenospheres: synthesis, characterization, and photocatalytic
ability under visible light, Appl. Surf. Sci. 258 (2012) 99899996.
[46] G. Wang, J. Li, L. Wang, X. Sun, J. Huang, Adsorption of dye from wastewater by
zeolites synthesized from y ash: kinetic and equilibrium studies, Chin. J.
Chem. Eng. 17 (3) (2009) 513521.
[47] D. Wu, P. Huo, Z. Lu, X. Gao, X. Liu, W. Shi, Y. Yan, Preparation of heteropolyacid/
TiO2/y-ash-cenosphere photocatalyst for the degradation of ciprooxacin
from aqueous solutions, Appl. Surf. Sci. 258 (2012) 70087015.
[48] D. Wu, B. Zhang, L. Yan, H. Kong, X. Wang, Effect of some additives on synthesis
of zeolite from coal y ash, Int. J. Miner. Process. 80 (2006) 266.
[49] R. Zallen, M.P. Moret, The optical absorption edge of brookite TiO2, Solid State
Commun. 137 (2006) 154157.
_
[50] M. Zenkiewicz,
Comparative study on the surface free energy of a solid
calculated by different methods, Polym. Test. 26 (2007) 1419.

Você também pode gostar