35
Subscriber access provided by Georgetown University | Lauinger and Blommer Libraries Industrial & Engineering Chemistry Research is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties. Article Kinetic and mechanism of enhanced photocatalytic activity under visible light using synthesized PrxCd1-xSe nanoparticles Alireza Khataee, Amirreza Khataee, Mehrangiz Fathinia, Younes Hanifehpour, and Sang Woo Joo Ind. Eng. Chem. Res., Just Accepted Manuscript • DOI: 10.1021/ie402352g • Publication Date (Web): 20 Aug 2013 Downloaded from http://pubs.acs.org on September 1, 2013 Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Kinetics and Mechanism of Enhanced Photocatalytic Activity under Visible Light Using Synthesized Pr x Cd 1– x Se Nanoparticles

Embed Size (px)

Citation preview

Subscriber access provided by Georgetown University | Lauinger and Blommer Libraries

Industrial & Engineering Chemistry Research is published by the American ChemicalSociety. 1155 Sixteenth Street N.W., Washington, DC 20036Published by American Chemical Society. Copyright © American Chemical Society.However, no copyright claim is made to original U.S. Government works, or worksproduced by employees of any Commonwealth realm Crown government in the courseof their duties.

Article

Kinetic and mechanism of enhanced photocatalytic activityunder visible light using synthesized PrxCd1-xSe nanoparticles

Alireza Khataee, Amirreza Khataee, Mehrangiz Fathinia, Younes Hanifehpour, and Sang Woo JooInd. Eng. Chem. Res., Just Accepted Manuscript • DOI: 10.1021/ie402352g • Publication Date (Web): 20 Aug 2013

Downloaded from http://pubs.acs.org on September 1, 2013

Just Accepted

“Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are postedonline prior to technical editing, formatting for publication and author proofing. The American ChemicalSociety provides “Just Accepted” as a free service to the research community to expedite thedissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscriptsappear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have beenfully peer reviewed, but should not be considered the official version of record. They are accessible to allreaders and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offeredto authors. Therefore, the “Just Accepted” Web site may not include all articles that will be publishedin the journal. After a manuscript is technically edited and formatted, it will be removed from the “JustAccepted” Web site and published as an ASAP article. Note that technical editing may introduce minorchanges to the manuscript text and/or graphics which could affect content, and all legal disclaimersand ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errorsor consequences arising from the use of information contained in these “Just Accepted” manuscripts.

2

Abstract

Praseodymium (Pr)-doped CdSe nanoparticles with different Pr contents were successfully

synthesized by an easy hydrothermal method. The prepared samples were characterized by X-

ray diffraction (XRD) and scanning electron microscopy (SEM). XRD analysis demonstrated

that the particles were excellently crystallized and corresponds to cubic CdSe phase. SEM

images indicated that the size of the particles were in the range of 50–100 nm. Photocatalytic

efficiency of pure and Pr-doped CdSe samples was evaluated by monitoring the decolorization

of malachite green (MG) in aqueous solution under visible light irradiation. The color removal

efficiency of 47.16 % for CdSe and 94.32 % for Pr0.06Cd0.94Se was achieved at 120 min. It was

found that the presence of the Cl−, SO4

−2 and HCO3

− ions obstructed the photocatalytic

degradation of MG, but the presence of S2O82−

accelerated it. Also, the effect of operational

parameters at optimum conditions was modeled regarding the nonlinear regression analysis.

Keywords: Photocatalysis; CdSe nanoparticles; Praseodymium; Decolorization; Non-

linear regression.

Page 1 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

3

1. Introduction

In most of industrial factories, including the textile, cosmetic, leather, pharmaceutical, and

food industries, large amounts of dyes are annually produced and used in many different

sections.1 The discharge of these huge amounts of pollutants including synthetic dyes to the

environment leads to chronic pollution of water sources because of their bioaccumulation in

aquatic and wildlife.2 Also, various problems associated to their carcinogenicity and toxicity

to aquatic life could be immerged.

Malachite green is an example of most commonly used dye for the dyeing of cotton, silk,

paper and leather between all other dyes of its category. Moreover, it is used as a pesticide in

aquaculture industry due to its efficiency in treatment of parasitic, fungal and bacterial

infections in fish and fish eggs.3 Despite the wide application ranges which were reported for

MG, its hazardous and carcinogenic effects limit its effective application. In this context, MG

can cause allergy, dermatitis, and cancer in humans.4 Additionally, it decreases food intake

capacity, growth and fertility rates, causes damage to heart, eyes, liver and kidney thereby

compromising aquatic life.3 Therefore, the development of new techniques for managing the

effluents containing such dye is crucial due to its harmful impacts on receiving waters.

Among various physical, chemical and biological techniques, heterogeneous photocatalysis

via the use of semiconductor photocatalysts has been considered as a cost-effective alternative

method for polluted water treatment. The advantages of photocatalytic technique over the

traditional ones are the rapid oxidation rate, high efficiency, no formation of polycyclic products

and oxidation of pollutants even in the low levels.2

Different semiconductors have been used to exhibit photocatalytic behavior. TiO2 and ZnO

are the most commonly utilized semiconductors for organic pollutant degradation.5, 6

However,

because of the wide band gap of TiO2 and ZnO (band gap around 3.2 eV and 3.37 eV for TiO2

and ZnO, respectively), their practical application is confined due to the need of an ultraviolet

Page 2 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

4

excitation source, which is only a small part (3−5%) of solar light.6-8

Therefore, to overcome this

problem, visible-light-driven photocatalysts were developed for photocatalysis application under

solar energy.9-12

Among these materials, small band gap semiconductors such as CdSe, the II–

VI semiconducting materials and the member of cadmium chalcogenide family, have aroused

extensive attention.13

Due to its low band gap and the rapid generation of electron-hole pairs

(charge carriers) by photoexcitation, CdSe has been regarded as an efficient semiconductor

under visible light irradiation.14, 15

However, still some issues are remained with the visible light activated semiconductors

such as CdSe which prevent their practical application: instability upon light illumination leads

to photocorrosion or photodissolution on a photocatalyst surface during photocatalytic reaction;

low adsorption capacity and separation efficiency of electron-hole pairs resulted in low

photocatalytic activity.16, 17

Generally, the adsorption capacity of CdSe particles could be improved by producing

CdSe nanomaterials containing enhanced specific surface area or depositing CdSe onto the

materials with high surface area. By preparing a hybrid or composite semiconductor materials

containing CdSe and metal ion doping of CdSe, the fast recombination of electron-hole pairs

could be efficiently inhibited and nanoparticles of photocatalysts are obtained. Regarding this

context, CdSe-Pt nanorods,18

hybrid CdSe-Au nanostructures,19

carbon nanomaterials-CdSe

composites20

have been successfully prepared and reported for degradation of various organic

compounds. However, according to the literature reports,21-24

metal ion doping of

semiconductors has considerably increased the photocatalytic efficiency of these materials. Yang

et al.17

reported that metal ion doping of semiconductors not only could affect their photo-

physical behavior, but also expedites the photochemical reaction. Moreover, metal ion may

change surface characteristics by the generation of a Schottky barrier through the metal in contact

with the semiconductor surface, which acts as an electron trap. There are also many rare earth

Page 3 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

5

metals such as Pr3+

,25

Sm3+

,23

Nd3+

,23

Ce3+

,26

Y3+ 27

and Eu3+

28

that have been utilized to promote

the separation of photogenerated electron and hole for improving photocatalytic property.

Nevertheless, there have been no reports regarding the improvement of CdSe photocatalytic

property under visible light irradiation via doping of transition rare earth metals such as Pr, a

lanthanide metal.

On the other hand, for the fabrication of CdSe samples, various methods such as the sol–

gel and electrochemical were wildly utilized.10, 29, 30

Most of these synthesis processes

generally require highly sophisticated equipment to produce high purity samples. In the case

of sol-gel technique, such high calcination temperatures will increase the nanoparticle size

and decrease the specific surface area.21

In order to produce highly photoactive

nanomaterials, a practical technique with a low temperature procedure should be applied to

synthesis samples with high surface area. The hydrothermal technique is a one-pot chemical

method, which facilitates the control of particle size, particle morphology and phase

composition by adjusting experimental parameters of the solution.31, 32

So, in this work we present a new method to prepare cubic CdSe nanoparticles using

cadmium acetate as a cationic source and sodium selenite as an anionic source in the presence

of ethylenediaminetetraacetic acid (EDTA), as template agent to avoid further aggregation of

synthesized nanoparticles, using a hydrothermal method. Pr was specifically selected here

due to its efficient function in improvement of TiO2 photocatalytic properties under visible

light in the previously reported papers.25, 33

It was reported that doping of Pr3+

to TiO2

effectively enhanced its photocatalytic activity under visible light through increasing the

separation efficiency of interfacial charges by Pr3+

and the red shift of the optical absorption

edge of TiO2 by Pr3+

doping.25

The physical properties of the synthesized samples were

characterized by XRD and SEM techniques. The prepared CdSe and PrxCd1-xSe nanoparticles

were used as a photocatalyst for decolorization of MG as a model pollutant (See Table 1).

Page 4 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

6

Table 1. Characteristics of Malachite Green.

Mw

(g/mol)

Color

index

number

maxλ

(nm)

Molecular

formula Chemical structure

Color index

name

364.911 42000 619 C23H25ClN2

C.I. Basic

Green 4

Moreover, the performance of synthesized samples in terms of decolorization efficiency

and kinetic rate constant were studied and compared. To the best of our knowledge, there is

no previous literature report concerning the use of CdSe and PrxCd1-xSe nanoparticles for the

removal of MG. Other objectives of this work are to investigate the effect of inorganic ions

on the decolorization efficiency of MG. Also, we have modeled the kinetic of the process

using nonlinear regression analysis.

2. Experimental details

2.1. Materials

All chemicals used in this study were of analytical grade and were used without further

purification. Cadmium acetate dihydrate (Cd(CH3COO)2.2H2O 98%), hydrazine

monohydrate (N2H4.H2O 80%), sodium selenite (Na2SeO3 99%) were purchased from Loba

Chemie Co. (India). NaOH, NaCl, NaHCO3 and K2S2O8 were purchased from

Merck.Praseodymium (III) nitrate hexahydrate (Pr(NO3)3.6H2O 99.99%), was purchased

from Aldrich. EDTA (C10H14N2O8Na2.2H2O) was purchased from Rankem Co. (India). MG

and ethanol (90%) was purchased from ShimiBoyakhsaz Co. (Iran) and Iran Daru Co. (Iran),

respectively.

Page 5 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

7

2.2. Synthesis of CdSe and Pr-doped CdSe samples

CdSe and Pr-doped CdSe nanoparticles with variable Pr mole fractions (0-10 % mol)

were prepared by hydrothermal method using N2H4.H2O as the reducing agent. In a typical

synthesis, 1 mmol Na2SeO3 powder and 1 mmol NaOH and appropriate molar ratios of

Pr(NO3)3.6H2O and Cd(CH3COO)2.2H2O were first dissolved in 70 mL distilled water. Under

middle speed stirring, 1 mmol EDTA was dissolved in 20 mL distilled water and added to the

above solution. After being stirred uniformly, hydrazine was added drop wise to the above

solution. The resulting solution was moved into a 150 mL teflon-lined stainless-steel

autoclave, placed in an oven at 180 ◦C for 24 h, and then the autoclave was allowed to cool to

room temperature naturally. As-synthesized CdSe and PrxCd1-xSe nanoparticles were

collected and washed with distilled water and ethanol several times in order to remove

residual impurities and then dried at 60 °C for 5 h. The final black powders were obtained as

a result.

2.3. Characterization instruments

To determine the crystal phase composition of the prepared CdSe and Pr-doped CdSe

samples, XRD characterization was carried out at room temperature using a D8 Advance,

Bruker, Germany diffractometer with monochromatic high-intensity CuKa radiation

(l=1.5406 A˚), the accelerating voltage of 40 kV and the emission current of 30 mA. The

Debye–Scherrer formula was employed to calculate the average crystalline size of the

catalysts.34

SEM (S-4200, Hitachi, Japan) was used to observe the surface state and

morphology of the prepared nanoparticles using an electron microscope.

Page 6 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

8

2.4. Photocatalytic activity and experimental procedures

The photocatalytic activity of undoped and Pr-doped CdSe nanoparticles was evaluated

by the decolorization of MG in an aqueous solution under visible light. In a typical process,

0.1 g of the photocatalyst powder was added into 100 mL MG solution with an initial

concentration of 5 mg/L. The suspension of photocatalyst and MG was magnetically stirred

in a quartz photoreactor in the dark for 15 min to establish an adsorption/desorption

equilibrium of the dye. Then, the solution was irradiated by a 6W fluorescent visible lamp

(GK-140, China) as the light source for a set irradiation time. Visible light irradiation of the

reactor was performed for 15, 30, 45, 60, 90 and 120 min. Samples were withdrawn regularly

from the reactor, and dispersed powders were removed in a centrifuge. The color removal

was evaluated by determining its absorbance at λmax= 619 nm by using UV–Vis

spectrophotometer, Lightwave S2000 (England). The decolorization efficiency (DE (%)) was

expressed as the percentage ratio of decolorized dye concentration to that of the initial one.

In the experiment requiring inhibitors, different amounts of probe molecules (NaCl,

Na2SO4, NaHCO3 and K2S2O8) were added to the MG solution before the introduction of the

PrxCd1-xSe as the catalyst. The continuance of the procedure was identical for the

photocatalytic activity evaluation as mentioned above.

3. Results and discussion

3.1. Characterization of the synthesized samples

The XRD patterns of the CdSe and Pr0.06Cd0.94Se nanoparticles are depicted in Figure 1a

and 1b, respectively. The XRD diffraction peaks at 2θ of 25.4°, 42°, and 49.6° can be

classified to the (100), (002), (101), (112) and (211) plane reflections. It can be associated

with cubic crystal structure CdSe according to the standard powder diffraction data (JCPDS

65-2891 for CdSe, cubic).35, 36

Figure 1b shows that, the same typical peaks for CdSe are

Page 7 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

9

detected for Pr0.06Cd0.94Se sample indicating that there is no change in the crystal structure

upon Pr doping. No peak for impurities was detected, confirming that the applied

hydrothermal method in this study was successful in synthesizing the samples. Moreover, the

sharp diffraction peaks in the XRD spectra of the synthesized samples express that the

synthesized products are high crystalline. Calculating from the Debye–Scherrer formula,34

the

average size of the crystals was about 15 nm. In order to further clarify the size and shape of

the nanoparticles, SEM image was taken at different magnifications.

Figure 1. Powder X-ray diffraction pattern of (a): CdSe and (b): Pr0.06Cd0.94Se samples.

Figures 2 and 3 show the SEM microphotographs of the CdSe and Pr doped CdSe

samples, respectively. In Figure 2, uniform and spherical nanoparticles of about 50-100 nm in

Page 8 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

10

diameter with a little agglomeration can be seen. In Figure 3, very uniform and spherical-

shaped Pr doped CdSe nanoparticles can be observed. The diameter of these particles is

around 50-100 nm. These figures confirm that doping of Pr3+

into the structure of CdSe does

not change the morphology of CdSe nanoparticles.

(a)

(b)

Figure 2. (a) and (b): SEM images of CdSe nanoparticles synthesized at 150 ºC and 24 h at

different magnifications.

Page 9 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

11

(a)

(b)

Figure 3. (a) and (b): SEM images of Pr0.06Cd0.94Se nanoparticles synthesized at 150 ºC and 24

h at different magnifications.

3.2. Effect of operating conditions on the photocatalysis of MG

3.2.1. Effect of Pr3+

content of PrxCd1-xSenanoparticles

To explore the optimum conditions of photocatalytic activity, the degradation process of MG

was studied in the presence of PrxCd1-xSe with different mole fraction (x = 0.00, 0.02, 0.06,

Page 10 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

12

0.08 and 0.10) under visible light irradiation. Figure 4 shows the decolorization efficiency of

MG over different Pr-doped CdSe photocatalysts during 120 min of reaction. From Figure 4a, it

can be clearly seen that the samples doped with appropriate amount of Pr ion had much higher

photocatalytic activity compared with pure CdSe, especially the sample with 6% molar ratio of

Pr, which exhibited the best photocatalytic activity. The reason of high photocatalytic activity of

Pr0.06Cd0.94Se can be explained by the following mechanism. Generally, rare earth metal ions

such as Pr3+

, can perform either as a mediator of interfacial charge transfer or as a recombination

center in the crystalline structure of CdSe.28, 37

So depending on the dopant concentration, the

doped CdSe activity or efficiency is changeable. At low mole fractions of dopant, Pr3+

ions can

capture photoinduced electrons retarding electron/hole recombination rate and thereafter

enhancing the interfacial charge transfer for MG degradation.38

However, when the mole

fractions of dopant is higher than the optimal value, the recombination rate may increase since

the distance between trapping sites in a crystal structure of CdSe is decreased which lead to

decrease in photocatalytic activity.21

So, an optimum content of Pr3+

is essential to separate

photoinduced electron/hole pairs and increase the life time of charge carries.

To understand the reaction kinetic of MG decolorization in this study, the following pseudo-

first-order kinetic equation is used to fit experimental data:

tk)C

Cln( app

0 = (1)

where kapparent (kapp) is the pseudo-first-order rate constant (min-1

), C0 is the initial concentration at

time 0 min, and C is the concentration at the reaction time of t min.

Page 11 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

13

(a)

(b)

(c)

Figure 4. (a): The effect of Pr3+

dopant content on the decolorization of 5 mg/L MG (catalyst

loading 1.0 g/L); (b): Apparent pseudo first-order reaction kinetic and (c): Variation of apparent

kinetic constant for MG decolorization during the photocatalytic reaction at different Pr3+

contents.

Page 12 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

14

The fitting curves of ln(C0/C) verses time during 0–120 min at different Pr3+

mole fractions

were plotted and presented in Figure 4b. Figure 4b indicates that photodecolorization of MG over

PrxCd1-xSesamples follows pseudo-first-order kinetic. Figure 4c shows the kinetic constant

during the photocatalytic reaction at different Pr3+

mole fractions by the pseudo first-order model.

As can be seen, the kinetic constant of photocatalytic reaction increases with increasing the Pr3+

content up to 6 % and then decreases. The possible reason for this behavior was explained above.

3.2.2. Enhanced photocatalytic decolorization mechanism of MG on Pr0.06Cd0.94Se

In order to properly illustrate the photocatalytic activity of the pure CdSe and the

Pr0.06Cd0.94Se samples and investigate the possible reaction mechanism, the changes in the

UV-Vis absorption spectra of MG during the photocatalytic process at different irradiation

times are shown in Figure 5. The decreasing concentration of MG during the photocatalytic

reaction is used to evaluate the activity of the photocatalysts. Two steps are included in the

photocatalytic removal of a dye: the adsorption of dye molecules, and the degradation. Figure 5a

and b represent both the adsorption and degradation effects for CdSe and Pr0.06Cd0.94Se samples

under visible light irradiation, respectively. At the identical conditions, the Pr0.06Cd0.94Se exhibits

significantly higher photocatalytic activity than that of pure CdSe. From Figure 5a, it can be

clearly seen that when pure CdSe used as a photocatalyst, MG concentration reaches to a constant

amount (47.16 % color removal efficiency) after 120 min of reaction and the degradation is

stopped. There are several probable reasons for the low activity of pure CdSe including

photocorrosion and photodissolution occurred on the surface of pure CdSe in the photocatalytic

reaction.17

The other major issue is that, when the band-gap energy of a semiconductor is low,

the life time of the produced electrons and holes in its surface is low.3 So, in the case of CdSe

the presumed reason is that, after a 60 min of irradiation and producing charge carries on the

surface of CdSe, the rate of recombination of charge carriers is dominated to the rate of other

Page 13 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

15

reactions in the solution, which lead to the decrease in its photocatalytic activity. However, in

Figure 5b, it is obvious that after 15 min of adsorption step and 120 min of irradiation, the

concentration of the MG solution gradually reduces with increasing irradiation time.

(a)

(b)

Figure 5. Adsorption and degradation of MG under visible light irradiation using (a): pure

CdSe and (b): Pr0.06Cd0.94Se nanoparticles

Therefore, these results indicate that Pr3+

dopant really could play an important role in the

improvement of photocatalytic activity of CdSe. This behavior can be attributed to the fact that,

Page 14 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

16

the number of the produced electrons and holes and their life time are dependent on the kind,

amount of dopant and size of the semiconductor particles.16, 38

Herein, the Pr3+

dopant

operates as an electron scavenger on the surface of CdSe, suppressing the recombination of

electron-hole pairs and enhancing their life time; so, the photocatalytic activity of

photocatalyst is increased.

Incorporating our achieved results into the other literature reports,5, 33, 39

the probable

mechanism for the enhanced photocatalysis of Pr0.06Cd0.94Se was proposed as follow. Under

the visible light irradiation, the electrons (e−) are excited from the valence band to the

conduction band of CdSe, and the holes (h+) are produced. Pr

3+ dopant in CdSe can

effectively scavenge the e− and inhibit their recombination with h

+ due to the existence of

partially filled 4f-orbital, 4f2. The produced Pr

2+ ions, with three electrons in 4f orbital, are

very unstable, so that the e− can be easily descavenged and passed to the adsorbed O2

molecules, promoting the O2•−

and •OH formation subsequently. Simultaneously, the

adsorbed photogenerated holes oxidize water molecules or surface bound hydroxide species

to generate •OH species. The degradation mechanism for the Pr0.06Cd0.94Se was shown in Eqs

2-11.

Page 15 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

17

)11(OHOHh

)10(HOHOHh

)9(OHOHeOH

)8(OHeHOOH

)7(OOHHO

)6()stepngtransferrielectron(OPrOPr

)5()stepscavengingelectron(PrPrO

)4(OeO

)3()stepscavengingelectron(PrePr

)2()he(SCdPrlightvisible SeCdPr

adsadsads

adsadsads2ads

adsadsadsads22

ads22adsadsads

adsadsads2

ads2

3

ads2

2

23ads2

ads2adsads2

2ads

3

adsads40.90.0640.90.06

•−+

+•+

−•−

−+•

•+−•

−•++

++−•

−•−

+−+

+−

→+

+→+

+→+

→++

→+

+→+

→+

→+

→+

+→+

Thus, the scavenging and transferring of the charge carriers in CdSe surface can be

attributed to Pr3+

dopant in its crystal structure which further improves the photocatalytic

activity of the catalyst.

3.2.3. Effect of photocatalyst concentration and reusability

The initial rate of photocatalytic decolorization process is dependent on the photocatalyst

concentration.21, 29, 36

In order to compare the effect of different dosage of Pr0.06Cd0.94Se (0.05–1.5

g/L) on the decolorization efficiency of MG, a series of experiments were done with 5 mg/L

initial concentration of MG. The results are presented in Figure 6a. Figure 6b indicates that

photocatalytic decolorization of MG with different amounts of Pr0.06Cd0.94Se samples follows

pseudo-first-order kinetic. Also, Figure 6c shows the kinetic constant during the photocatalytic

reaction using different amounts of Pr0.06Cd0.94Se by the pseudo first-order model. As can be

seen in Figure 6c, the kapp significantly increases with an increase in the amount of photocatalyst

from 0.05 to 1.0 g/L; it reaches to the maximum value (0.0223 min-1

) with the content of 1.0 g/L.

By further increase of photocatalyst amount, the kapp is decreased slightly. These results indicate

Page 16 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

18

that the photocatalyst dosage of 1.0 g/L would be the optimum amount. The assumed reason is

that, when the concentration of photocatalyst is increased, the available photocatalyst active sites,

the number of photons adsorbed and the number of dye molecules adsorbed are also increased,

so, the decolorization efficiency is increased in the presence of 1.0 g/L photocatalyst.16, 21

However, by further increase of the photocatalyst dosage beyond the optimum amount, the kapp is

diminished because of increasing the turbidity of the suspension. So, the penetration depth of the

photons reduced and less photocatalyst nanoparticles could be excited.40

This leads to a decrease

in the color removal efficiency.

(a)

(b)

Page 17 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

19

(c)

Figure 6. (a): The effect of photocatalyst loading on the decolorization of 5 mg/L MG by the

Pr0.06Cd0.94Se; (b): Apparent pseudo first-order reaction kinetics and (c): Variation of apparent

kinetic constant for MG decolorization during the photocatalytic reaction at different

photocatalyst concentration.

The reusability is one of the most important factors for a photocatalyst. Figure 7 shows the

reusability tests of Pr0.06Cd0.94Se photocatalyst in the decolorization of MG, during 10 cycle

experiments under optimum conditions as: 5 mg/L of MG, 1.0 g/L of Pr0.06Cd0.94Se photocatalyst

and irradiation time of 120 min. After each decolorization experiment, the photocatalyst was

washed with distilled water and then dried at 60 °C for 5 h and then used in new experiment.

As can be seen in Figure 7, Pr0.06Cd0.94Se exhibited excellent chemical stability without any

considerable decomposition or photocorrosion during the 10 cycles of photocatalytic reaction

which is an important advantageous for practical applications.

Page 18 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

20

Figure 7. Reusability behavior of Pr0.06Cd0.94Se in photocatalytic decolorization of 5 mg/L of MG

in the presence of 1.0 g/L of Pr0.06Cd0.94Se and irradiation time of 120 min.

3.2.4. Effect of initial MG concentration

Photocatalytic decolorization of MG was investigated at different concentrations of dye

solutions varying from 5 to 15 mg/L, in the presence of 1.0 g/L Pr0.06Cd0.94Se under UV

irradiation. Figure 8a shows the decolorization efficiency of different concentrations of MG dye

solution in the certain irradiation time intervals. Figure 8b shows that photocatalytic

decolorization of MG with its various initial concentrations follows pseudo-first-order kinetic.

Also, Figure 8c presents the kinetic constant during the photocatalytic reaction using different

initial concentrations of MG. It can be observed from Figure 8c that the kapp declines with

increasing the initial concentration of the dye.

This behavior can be related to the several factors. By increasing the dye concentration,

more and more molecules of the dye get adsorbed on the surface of the photocatalyst.

Therefore, requirement of the reactive oxygen species (such as •OH and •O2

−) in order to

degrade the dye molecules, increases. However, the formation of •OH and •O2

− on the

photocatalyst surface remains constant for a given light intensity, catalyst dosage and

Page 19 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

21

duration of irradiation. Hence, the available hydroxyl radicals were not enough to oxidize the

high concentration of MG and more generated intermediates from its degradation.2

Accordingly, the degradation efficiency of the dye decreases as the concentration increases.

Moreover, the increasing of substrate concentration can lead to the production of intermediates,

which may adsorb on the surface of the catalyst. Slow diffusion of the produced

intermediates from the catalyst surface can result in the deactivation of active sites of the

photocatalyst and consequently, a reduction in the degradation efficiency. Another reason for

reducing the color removal at high substrate concentrations may be due to the absorption of light

photon by the dye itself that leads to a lesser availability of photons for hydroxyl radical

production.1, 41

(a)

Page 20 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

22

(b)

(c)

Figure 8. (a): The effect of MG concentration on the decolorization of 5 mg/L MG(Pr0.06Cd0.94Se

loading 1.0 g/L); (b): Apparent pseudo first-order reaction kinetics; and (c): Variation of apparent

kinetic constant for MG decolorization during the photocatalytic reaction at different MG

concentration.

3.2.5. Effect of inorganic ions on the photocatalytic activity

The effect of different inorganic ions including NaCl, Na2SO4, NaHCO3 and K2S2O8were

tested on the kapp of photocatalytic reaction. The decolorization efficiency of MG in the presence

of inorganic ions such as Cl−, SO4

2−, HCO3

− is shown in Figure 9a. In the absence of inorganic

Page 21 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

23

ions, the adsorption efficiency during 15 min of reaction under dark was 35.59 % and the

decolorization efficiency of MG during 120 min of photocatalysis was 94.32 %. Figure 9a shows

that, when the NaCl, Na2SO4 and NaHCO3are added to the reaction solution separately, with the

concentration ratio of 1:1 to MG, both the adsorption and decolorization efficiency of MG

decrease considerably. This behavior can be related to the adsorption properties of these ions.

According to Figure 9a, the adsorption of Cl−, SO4

2− and HCO3

− ions on the Pr0.06Cd0.94Se surface

is stronger than that of MG. So, the surface of the Pr0.06Cd0.94Se is dominated by these ions

which lead to a decrease in adsorption efficiency of MG. This phenomenon demonstrates that the

preferential adsorption of pollutant molecules on the surface of the catalyst is a critical step in

photocatalytic reaction.42

Moreover, Figure 9 shows that the addition of the three mentioned inorganic anions decreases

the photocatalytic decolorization efficiency of MG. The inhibition effect of these ions is due to

the h+ and

•OH scavenging properties of these adsorbed ions as shown in Eqs. 12–17. The results

are in good agreement with the published papers which reported a strong inhibiting effect of

inorganic ions.42, 43

)17(OHCOOHCO

)16(OHCOOHHCO

)15(OHSOOHSO

)14(SOhSO

)13(ClClCl

)12(ClhCl

adsads3adsads2

3

2ads3adsads3

adsads4adsads2

4

ads4adsads2

4

ads2adsads

adsadsads

−−••−

−••−

−−••−

−•+−

−•−•

•+−

+→+

+→+

+→+

→+

→+

→+

Moreover, the above reactions demonstrate that the CdSe surface active sites are blocked by

the produced radical anions which are not easily oxidizable, and subsequently the photocatalytic

activity decreases. Consequently, it can be deduced that MG is first adsorbed on the surface of the

catalyst and then is oxidized under the effect of holes and hydroxyl radicals.

Page 22 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

24

(a)

(b)

(c)

Figure 9. (a):The effect of addition of Cl−, SO4

2− and HCO3

−ions on the decolorization of 5 mg/L

MG (Pr0.06Cd0.94Se loading 1.0 g/L); (b): Apparent pseudo first-order reaction kinetics; and (c):

Variation of apparent kinetic constant in the presence of different inorganic ions with the

concentration ratio of 1:1 to MG.

Page 23 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

25

The effect of the addition of K2S2O8 on the photocatalytic activity is shown in Figure 10.

With the addition of K2S2O8, the photocatalytic activity is significantly increased. When the ratio

of K2S2O8 to MG was 6:1, the decolorization efficiency of MG is 98.25 % after being irradiated

for 60 min. The persulfate ions (S2O82−

) are strong electron acceptor and they can capture the

photogenerated electrons rapidly to produce hydroxyl radicals through Eqs. 18 and 19.42

)19(HOHSOOHSO

)18(SOSOeOS

adsads2

4ads2ads4

ads4ads2

4adsads2

82

+•−−•

−•−−−

++→+

+→+

According to the number of previous studies,42, 44

the persulfate ion is a more efficient

electron acceptor than molecular oxygen. So the addition of K2S2O8 decreased the recombination

of electrons and holes, thus greatly enhanced the decolorization efficiency. This results

reconfirmed the critical rule of •OH radicals in the degradation of MG using Pr0.06Cd0.94Se

nanocatalyst under visible light irradiation.

(a)

Page 24 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

26

(b)

(c)

Figure 10. (a):The effect of addition of S2O82−

ions, with different the concentration ratios to

MG, on the decolorization of 5 mg/L MG(Pr0.06Cd0.94Se loading 1.0 g/L); (b): Apparent pseudo

first-order reaction kinetics; and (c): Variation of apparent kinetic constant in the presence of

different concentrations of S2O82−

ions.

3.2.6. Kinetic modeling of the photocatalytic process

The aim of this section is to develop a kinetic model to estimate the pseudo-first order rate

constant (kapp) of MG degradation by photocatalytic process using Pr0.06Cd0.94Se nanocatalyst

under visible light irradiation. Non-linear regression analysis was used to find the relation

between kapp and operational parameters including initial concentration of MG, the presence

of different inorganic inhibitors with the concentration ratio of 1:1 to MG and initial

Page 25 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

27

concentration of K2S2O8 in the optimum concentration of catalyst dosage of 0.1 g/L.

According to pseudo-first order kinetic assumption, when straight lines with high correlation

coefficient values (R2> 0.90) were obtained from the plots of ln(C0/C) versus process time,

the proposed kinetic was confirmed.45

Moreover, the non-linear relation of operational variables and kapp can be also modeled

with empirical power-law type separately as Eq. 20.46

kapp= α (operational variable)β (20)

Where α and β are the model parameters.

So, the non-linear relation of operational variables including initial MG concentration, the

presence of different inorganic inhibitors with the concentration ratio of 1:1 to MG and initial

concentration of K2S2O8 and kapp were modeled with empirical power-law separately. The

model parameters (α and β) were calculated for each operational variable with non-linear

regression analysis of experimental data, and the obtained results are shown in Figures 8c, 9c

and 10c. The kapp declines by increasing of initial concentration of MG (Figure 8c) and also it

is reduced in the presence of various inorganic inhibitors (Figure 9c). On the other hand, the

kapp increases by increasing the initial concentration of K2S2O8 (Figure 10c).

As it was discussed in previous section, degradation of MG follows pseudo-first order

kinetic, and it is deduced that the kapp is a function of operational variables as follows:

[ ]

[ ] cb

0

a

0822'

app]Inhibitor[MG

OSKkk = (21)

Non-linear analysis regression was employed to estimate a, b, and c. Then, with known

values of kapp, [K2S2O8]0, [MG]0 and [Inhibitor],which were applied in experimental

conditions, k' can be computed for each run. The average values of k' was attained as 0.0458

(mg/L min). By substituting the estimated values into Eq. (21), the Eq. 22 was obtained for

prediction of kapp at various experimental conditions:

Page 26 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

28

[ ]

[ ] 795.0740.0

0

704.0

0822

app]Inhibitor[MG

OSK0458.0k = (22)

A comparison between experimental and calculated kapp, which was obtained from the

mathematical model (Eq. 22), was exhibited in Figure 11. It can be perceived that the

calculated results from the kinetic model are in good agreement with experimental data (R2=

0.999).

Figure 11. A comparison between experimental and calculated apparent reaction rate constants

obtained from expression 22, the new kinetic model.

4. Conclusions

In this research, a simple and efficient hydrothermal method was applied for the synthesis of

CdSe and Pr doped CdSe nanoparticles as a new visible light activated photocatalyst. The results

of XRD and SEM analyses confirmed that the doping of Pr3+

ions into CdSe structure neither

change the crystal structure nor the morphology of CdSe nanoparticles. Synthesized CdSe and

Pr doped CdSe nanoparticles were used as the photocatalyst for removal of MG under visible

light irradiation. The decolorization efficiency was 47.16 % and 94.32 % for CdSe and

Pr0.06Cd0.94Se, respectively at 120 min of photocatalytic process. The mole fraction of 0.06 of Pr

in nanoparticles showed the highest photocatalytic activity. It was shown that the presence of the

Page 27 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

29

inorganic ions such as Cl−, SO4

2− and HCO3

− reduced the decolorization efficiency of MG, but

the presence of S2O82−

accelerated the reaction. Furthermore, non-linear regression analysis was

applied to develop a kinetic model for estimating the kapp of photocatalytic decolorization. The

experimental results for kapp were in good agreement with theoretically calculated data (R2=

0.999) and confirmed the significance of the mathematical model.

Acknowledgments

The authors thank the University of Tabriz, Iran for all the supports provided. This work was

funded by the Grant 2011-0014246 of the National Research Foundation of South Korea.

References

(1) Fathinia, M.; Khataee, A. R.; Zarei, M.; Aber, S., Comparative photocatalytic degradation

of two dyes on immobilized TiO2 nanoparticles: Effect of dye molecular structure and

response surface approach. J. Molecul. Catal. A-Chem. 2010, 333, 73.

(2) Khataee, A. R.; Fathinia, M.; Aber, S.; Zarei, M., Optimization of photocatalytic

treatment of dye solution on supported TiO2 nanoparticles by central composite design:

Intermediates identification. J. Hazard. Mater. 2010, 181, 886.

(3) Rajabi, H. R.; Khani, O.; Shamsipur, M.; Vatanpour, V., High-performance pure and

Fe3+

-ion doped ZnS quantum dots as green nanophotocatalysts for the removal of malachite

green under UV-light irradiation. J. Hazard. Mater. 2013, 250, 370.

(4) Ahmad, M. A.; Alrozi, R., Removal of malachite green dye from aqueous solution using

rambutan peel-based activated carbon: Equilibrium, kinetic and thermodynamic studies.

Chem. Eng. J. 2011, 171, 510.

Page 28 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

30

(5) Sin, J.-C.; Lam, S.-M.; Lee, K.-T.; Mohamed, A. R., Photocatalytic performance of novel

samarium-doped spherical-like ZnO hierarchical nanostructures under visible light irradiation

for 2,4-dichlorophenol degradation. J. Colloid Interface Sci. 2013, 401, 40.

(6) Daghrir, R.; Drogui, P.; Robert, D., Modified TiO2 for environmental photocatalytic

applications: A review. Ind. Eng. Chem. Res. 2013, 52, 3581.

(7) Khataee, A. R.; Fathinia, M.; Aber, S., Kinetic modeling of liquid phase photocatalysis on

supported TiO2 nanoparticles in a rectangular flat-plate photoreactor. Ind. Eng. Chem. Res.

2010, 49, 12358.

(8) Riaz, N.; Chong, F. K.; Man, Z. B.; Khan, M. S.; Dutta, B. K., Photodegradation of

orange II under visible light using Cu–Ni/TiO2: influence of Cu:Ni mass composition,

preparation, and calcination temperature. Ind. Eng. Chem. Res. 2013, 52, 4491.

(9) Wu, Y.; Xu, F.; Guo, D.; Gao, Z.; Wu, D.; Jiang, K., Synthesis of ZnO/CdSe hierarchical

heterostructure with improved visible photocatalytic efficiency. Appl. Surf. Sci. 2013, 274,

39.

(10) Ho, W.; Yu, J. C., Sonochemical synthesis and visible light photocatalytic behavior of

CdSe and CdSe/TiO2 nanoparticles. J. Molecul. Catal. A-Chem. 2006, 247, 268.

(11) Mahlambi, M. M.; Mishra, A. K.; Mishra, S. B.; Krause, R. W.; Mamba, B. B.; Raichur,

A. M., Effect of metal ions (Ag, Co, Ni, and Pd) on the visible light degradation of

Rhodamine B by carbon-covered alumina-supported TiO2 in aqueous solutions. Ind. Eng.

Chem. Res. 2013, 52, 1783.

(12) Xiong, P.; Wang, L.; Sun, X.; Xu, B.; Wang, X., Ternary titania–cobalt ferrite–

polyaniline nanocomposite: A magnetically recyclable hybrid for adsorption and

photodegradation of dyes under visible light. Ind. Eng. Chem. Res. 2013, 52, 10105.

(13) Behboudnia, M.; Azizianekalandaragh, Y., Synthesis and characterization of CdSe

semiconductor nanoparticles by ultrasonic irradiation. Mater. Sci. Eng. 2007, 138, 65.

Page 29 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

31

(14) Chen, M.-L.; Meng, Z.-D.; Zhu, L.; Park, C.-Y.; Choi, J.-G.; Ghosh, T.; Cho, K.-Y.; Oh,

W.-C., Synthesis of carbon nanomaterials-CdSe composites and their photocatalytic activity

for degradation of methylene blue. J. Nanomaterials 2012, 2012, 21.

(15) Robel, I.; Subramanian, V.; Kuno, M.; Kamat, P. V., Quantum dot solar cells. harvesting

light energy with CdSe nanocrystals molecularly linked to mesoscopic TiO2 films. J. Am.

Chem. Soc. 2006, 128, 2385.

(16) Xiao, Q.; Si, Z.; Zhang, J.; Xiao, C.; Tan, X., Photoinduced hydroxyl radical and

photocatalytic activity of samarium-doped TiO2 nanocrystalline. J. Hazard. Mater. 2008,

150, 62.

(17) Yang, F.; Yan, N.-N.; Huang, S.; Sun, Q.; Zhang, L.-Z.; Yu, Y., Zn-Doped CdS

Nanoarchitectures Prepared by Hydrothermal Synthesis: Mechanism for Enhanced

Photocatalytic Activity and Stability under Visible Light. J. Phys. Chem. C 2012, 116, 9078.

(18) Elmalem, E.; Saunders, A. E.; Costi, R.; Salant, A.; Banin, U., Growth of photocatalytic

CdSe–Pt nanorods and nanonets. Adv. Matter. 2008, 20, 4312.

(19) Costi, R.; Saunders, A. E.; Elmalem, E.; Salant, A.; Banin, U., Visible light-induced

charge retention and photocatalysis with hybrid CdSe-Au nanodumbbells. Nano Lett 2008, 8,

637.

(20) Meng, Z.-D.; Zhu, L.; Oh, W.-C., Preparation and high visible-light-induced

photocatalytic activity of CdSe and CdSe-C60 nanoparticles. J. Ind. Eng. Chem. 2012, 18,

2004.

(21) Asiltürk, M.; Sayılkan, F.; Arpaç, E., Effect of Fe3+

ion doping to TiO2 on the

photocatalytic degradation of Malachite Green dye under UV and vis-irradiation. J. Photoch.

Photobio. A-Chem. 2009, 203, 64.

(22) Kumar, S.; Kumari, N.; Singh, S.; Singh, T.; Jain, S., Doping studies of Tb (terbium) and

Cu (copper) on CdSe nanorods. Colloid. Surface. A 2011, 389, 1.

Page 30 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

32

(23) Liang, C.-H.; Li, F.-B.; Liu, C.-S.; Lü, J.-L.; Wang, X.-G., The enhancement of

adsorption and photocatalytic activity of rare earth ions doped TiO2 for the degradation of

Orange I. Dyes Pigments 2008, 76, 477.

(24) Liu, T.-x.; Li, X.-z.; Li, F.-b., Enhanced photocatalytic activity of Ce3+

-TiO2 hydrosols

in aqueous and gaseous phases. Chem. Eng. J. 2010, 157, 475.

(25) Liang, C.; Liu, C.; Li, F.; Wu, F., The effect of Praseodymium on the adsorption and

photocatalytic degradation of azo dye in aqueous Pr3+

-TiO2 suspension. Chem. Eng. J. 2009,

147, 219.

(26) Xiao, J.; Peng, T.; Li, R.; Peng, Z.; Yan, C., Preparation, phase transformation and

photocatalytic activities of cerium-doped mesoporous titania nanoparticles. J. Solid. State.

Chem. 2006, 179, 1161.

(27) Niu, X.; Li, S.; Chu, H.; Zhou, J., Preparation, characterization of Y3+

-doped TiO2

nanoparticles and their photocatalytic activities for methyl orange degradation. J. Rare Earth.

2011, 29, 225.

(28) Jose, G.; Amrutha, K. A.; Toney, T. F.; Thomas, V.; Joseph, C.; Ittyachen, M. A.;

Unnikrishnan, N. V., Structural and optical characterization of Eu3+

/CdSe nanocrystal

containing silica glass. Mater. Chem. Phys. 2006, 96, 381.

(29) Ghosh, T.; Lee, J.-H.; Meng, Z.-D.; Ullah, K.; Park, C.-Y.; Nikam, V.; Oh, W.-C.,

Graphene oxide based CdSe photocatalysts: Synthesis, characterization and comparative

photocatalytic efficiency of rhodamine B and industrial dye. Mater. Res. Bull. 2013, 48,

1268.

(30) Tian, L.; Ding, J.; Zhang, W.; Yang, H.; Fu, W.; Zhou, X.; Zhao, W.; Zhang, L.; Fan, X.,

Synthesis and photoelectric characterization of semiconductor CdSe microrod array by a

simple electrochemical synthesis method. Appl. Surf. Sci. 2011, 257, 10535.

Page 31 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

33

(31) Khataee, A. R.; Hosseini, M.; Hanifehpour, Y.; Safarpour, M.; Joo, S. W., Hydrothermal

synthesis and characterization of Nd-doped ZnSe nanoparticles with enhanced visible light

photocatalytic activity. Res. Chem. Intermed. 2012, 1.

(32) Williams, J. V.; Adams, C. N.; Kotov, N. A.; Savage, P. E., Hydrothermal synthesis of

CdSe nanoparticles. Ind. Eng. Chem. Res. 2007, 46, 4358.

(33) Chiou, C.-H.; Juang, R.-S., Photocatalytic degradation of phenol in aqueous solutions by

Pr-doped TiO2 nanoparticles. J. Hazard. Mater. 2007, 149, 1.

(34) Patterson, A. L., The Scherrer Formula for X-Ray Particle Size Determination. Physical

Review 1939, 56, 978.

(35) Raevskaya, A. E.; Stroyuk, A. L.; Kuchmiy, S. Y.; Azhniuk, Y. M.; Dzhagan, V. M.;

Yukhymchuk, V. O.; Valakh, M. Y., Growth and spectroscopic characterization of CdSe

nanoparticles synthesized from CdCl2 and Na2SeSO3 in aqueous gelatine solutions. Colloid.

Surface. A 2006, 290, 304.

(36) Ghosh, T.; Ullah, K.; Nikam, V.; Park, C.-Y.; Meng, Z.-D.; Oh, W.-C., The

characteristic study and sonocatalytic performance of CdSe–graphene as catalyst in the

degradation of azo dyes in aqueous solution under dark conditions. Ultrason. Sonochem.

2013, 20, 768.

(37) Jose, G.; Joseph, C.; Ittyachen, M. A.; Unnikrishnan, N. V., Structural and optical

characterization of CdSe nanocrystallites/rare earth ions in sol–gel glasses. Opt. Mater. 2007,

29, 1495.

(38) Ma, Y.; Zhang, J.; Tian, B.; Chen, F.; Wang, L., Synthesis and characterization of

thermally stable Sm, N co-doped TiO2 with highly visible light activity. J. Hazard. Mater.

2010, 182, 386.

Page 32 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960

34

(39) Wu, Y.; Xu, F.; Guo, D.; Gao, Z.; Wu, D.; Jiang, K., Synthesis of ZnO/CdSe

hierarchical heterostructure with improved visible photocatalytic efficiency. Appl. Surf. Sci.

2013, 274, 39.

(40) Avasarala, B. K.; Tirukkovalluri, S. R.; Bojja, S., Photocatalytic degradation of

monocrotophos pesticide-an endocrine disruptor by magnesium doped titania. J. Hazard.

Mater. 2011, 186, 1234.

(41) Damodar, R. A.; Jagannathan, K.; Swaminathan, T., Decolourization of reactive dyes by

thin film immobilized surface photoreactor using solar irradiation. Sol. Energy 2007, 81, 1.

(42) Ma, Y.; Zhang, J.; Tian, B.; Chen, F.; Bao, S.; Anpo, M., Synthesis of visible light-

driven Eu, N co-doped TiO2 and the mechanism of the degradation of salicylic acid. Res.

Chem. Intermed. 2012, 38, 1947.

(43) Abdullah, M.; Low, G. K. C.; Matthews, R. W., Effects of common inorganic anions on

rates of photocatalytic oxidation of organic carbon over illuminated titanium dioxide. J. Phys.

Chem. 1990, 94, 6820.

(44) Qourzal, S.; Barka, N.; Tamimi, M.; Assabbane, A.; Ait-Ichou, Y., Photodegradation of

2-naphthol in water by artificial light illumination using TiO2 photocatalyst: Identification of

intermediates and the reaction pathway. Appl. Catal. A-Gen. 2008, 334, 386.

(45) Devi, L. G.; Rajashekhar, K. E., A kinetic model based on non-linear regression analysis

is proposed for the degradation of phenol under UV/solar light using nitrogen doped TiO2. J.

Molecul. Catal. A-Chem. 2010, 334, 65.

(46) Marandi, R.; Olya, M. E.; Vahid, B.; Khosravi, M.; Hatami, M., Kinetic modeling of

photocatalytic degradation of an azo dye using nano-TiO2/polyester. Environ. Eng. Sci. 2012,

29, 957.

Page 33 of 35

ACS Paragon Plus Environment

Industrial & Engineering Chemistry Research

123456789101112131415161718192021222324252627282930313233343536373839404142434445464748495051525354555657585960