Вы находитесь на странице: 1из 16

Online Proofing System

1. Corrections should be marked with the Adobe Annotation & Comment Tools below:

2. To save your proof corrections, click the Publish Comments button.

Publishing your comments saves the marked up version of your proof


to a centralized location in Wileys Online Proofing System. Corrections
dont have to be marked in one sitting you can publish corrections
and log back in at a later time to add more.

3. When your proof review is complete we recommend


you download a copy of your annotated proof for
reference in any future correspondence concerning
the article before publication. You can do this by
clicking on the icon to the right of the Publish
Comments button and selecting Save as Archive
Copy.

4. When your proof review is complete and you are ready to


send corrections to the publisher click the Complete
Proof Review button that appears above the proof in
your web browser window. Do not click the Complete
Proof Review button without replying to any author
queries found on the last page of your proof. Incomplete
proof reviews will cause a delay in publication. Note:
Once you click Complete Proof Review you will not be
able to mark any further comments or corrections.

Firefox, Chrome, Safari Users

If your PDF article proof opens in any PDF viewer other than Adobe Reader or Adobe Acrobat, you will not be able to
mark corrections and query responses, nor save them. To mark and save corrections, please follow these instructions
to disable the built-in browser PDF viewers in Firefox, Chrome, and Safari so the PDF article proof opens in Adobe within
a Firefox or Chrome browser window.

USING e-ANNOTATION TOOLS FOR ELECTRONIC PROOF CORRECTION


Required software to e-Annotate PDFs: Adobe Acrobat Professional or Adobe Reader (version 8.0 or
above). (Note that this document uses screenshots from Adobe Reader X)
The latest version of Acrobat Reader can be downloaded for free at: http://get.adobe.com/reader/
Once you have Acrobat Reader open on your computer, click on the Comment tab at the right of the toolbar:

This will open up a panel down the right side of the document. The majority of
tools you will use for annotating your proof will be in the Annotations section,
pictured opposite. Weve picked out some of these tools below:

1. Replace (Ins) Tool for replacing text.

2. Strikethrough (Del) Tool for deleting text.

Strikes a line through text and opens up a text


box where replacement text can be entered.
How to use it

Strikes a red line through text that is to be


deleted.
How to use it

Highlight a word or sentence.

Highlight a word or sentence.

Click on the Replace (Ins) icon in the Annotations


section.

Click on the Strikethrough (Del) icon in the


Annotations section.

Type the replacement text into the blue box that


appears.

3. Add note to text Tool for highlighting a section


to be changed to bold or italic.

4. Add sticky note Tool for making notes at


specific points in the text.

Highlights text in yellow and opens up a text


box where comments can be entered.
How to use it

Marks a point in the proof where a comment


needs to be highlighted.
How to use it

Highlight the relevant section of text.

Click on the Add note to text icon in the


Annotations section.

Click on the Add sticky note icon in the


Annotations section.

Click at the point in the proof where the comment


should be inserted.

Type the comment into the yellow box that


appears.

Type instruction on what should be changed


regarding the text into the yellow box that
appears.

USING e-ANNOTATION TOOLS FOR ELECTRONIC PROOF CORRECTION

5. Attach File Tool for inserting large amounts of


text or replacement figures.

6. Add stamp Tool for approving a proof if no


corrections are required.

Inserts an icon linking to the attached file in the


appropriate pace in the text.
How to use it

Inserts a selected stamp onto an appropriate


place in the proof.
How to use it

Click on the Attach File icon in the Annotations


section.

Click on the Add stamp icon in the Annotations


section.

Click on the proof to where youd like the attached


file to be linked.

Select the file to be attached from your computer


or network.

Select the stamp you want to use. (The Approved


stamp is usually available directly in the menu that
appears).

Click on the proof where youd like the stamp to


appear. (Where a proof is to be approved as it is,
this would normally be on the first page).

Select the colour and type of icon that will appear


in the proof. Click OK.

7. Drawing Markups Tools for drawing shapes, lines and freeform


annotations on proofs and commenting on these marks.
Allows shapes, lines and freeform annotations to be drawn on proofs and for
comment to be made on these marks..

How to use it

Click on one of the shapes in the Drawing


Markups section.

Click on the proof at the relevant point and


draw the selected shape with the cursor.

To add a comment to the drawn shape,


move the cursor over the shape until an
arrowhead appears.

Double click on the shape and type any


text in the red box that appears.

For further information on how to annotate proofs, click on the Help menu to reveal a list of further options:

MSP No.

Dispatch: August 21, 2015

CE: Sharin

CJCE

22315

No. of Pages: 11

PE: Tom O'Brien

Adsorptive Removal of Dyes from Synthetic and Real Textile


Wastewater Using Magnetic Iron Oxide Nanoparticles:
Thermodynamic and Mechanistic Insights
Q1

Nashaat N. Nassar,1,2* Nedal N. Marei,1,2 Gerardo Vitale1 and Laith A. Arar2

1. Department of Chemical and Petroleum Engineering, University of Calgary, 2500 University Drive Northwest, Calgary, AB,
T2N 1N4, Canada
2. Department of Chemical Engineering, An-Najah National University, P.O. Box 7, Nablus, Palestine

Keywords: textile, adsorption, nanoparticles, iron oxide, wastewater, maghemite

INTRODUCTION

technology related to the application of materials at nanoscale,


1100 nm) has emerged as an alternative method to common
adsorbents in treating efuent wastewater.[9,1521] Hence, magnetic
nanoadsorbents could be employed to rapidly adsorb pollutants
from wastewater, and then be easily separated from the medium
using a simple magnetic eld. Recently, because of their unique and
unexpected chemical and physical properties compared to its
counterparts, iron-based nanoadsorbents have emerged as an
alternate source for conventional adsorbents. Properties such as
excellent magnetic behaviour, chemical stability, biocompatibility,
amphoteric surface activity, high adsorption capacity, enhanced
catalytic activity, and dispersability have been reported.[20,2230]
Further, several researchers have reported on the employment of
iron-based nanoadsorbents in removing various pollutants from
wastewater.[1921,23,31,32] However, to the best of our knowledge,
there have been no reports on the application of magnetic assistednanoparticle adsorption technology to dye removal from real textile
wastewater.
In this work, g-Fe2O3 (maghemite) nanoadsorbents are
employed for the rst time in dye removal from a real textile
wastewater solution. To understand the adsorption mechanisms,
two model dyes with known chemical structures were explored for
their adsorptive removal by g-Fe2O3 nanoadsorbents. Further, we

TE

EC

he textile industry is one of the traditional and most-widely


practiced industries in the Middle East. This industry, which
is a major consumer of fresh water, discharges signicant
amounts of colours and organic dyestuffs in its efuents.[1] The
presence of these compounds in wastewater presents an environmental concern because these compounds are not just toxic (even at
very low concentrations, < 1 mg/L), but also non-biodegradable,
thus, they will exist for a long time in the environment.[2] Therefore,
removing dye compounds from wastewater is essential before
discharging them into water bodies. Various dyes with complex
chemical structures primarily based on aromatic, heterocyclic, and
azo compounds, such as aromatic amine (C6H5-NH2), phenyl
(C6H5CH2), naphthyl (C10H7CH2), and azo (-N N-) groups, are
well-known in literature.[38] These compounds are hard to remove
or treat with conventional biological treatment methods.[9] A
number of conventional chemical and physical treatment methods
have been reported for decolourization of textile wastewater,
including coagulation-occulation, precipitation, advanced oxidation, ion exchange, membrane ltration, adsorption onto activated
carbon and low cost adsorbents, etc.[1,46,1012] However, these
conventional treatment methods suffer from high cost and inability
to meet permissible disposal levels. Sorption has been widely used
for the removal of different dyes and other hazardous species from
wastewater using common sorbents, including activated carbon,
zeolites, and those prepared from industrial solid byproducts.[13]
However, some common adsorbents suffer from low adsorption
afnity, selectivity, and capacity.[13,14] In addition, most of these
types of adsorbents show unsatisfactory regeneration and cycling
ability,[14] which limits their application on an industrial scale.
Magnetic assisted-nanoparticle adsorption technology (i.e., the

VOLUME 9999, 2015

PR

FS

Magnetic iron-oxide nanoparticles exhibit high efciency in wastewater treatment for many important reasons, including that they can remove
contaminants from wastewater rapidly owing to their high external surface area/unit mass and interstice reactivity. Additionally, this type of iron
oxide can easily be separated using a magnet after nishing the treatment process, can be used as a catalyst for the decomposition of adsorbed
contaminants and thus reduce sludge formation, and can cost-effectively meet the environmental regulations for wastewater treatment since it can
be prepared in situ where treatment is needed via various techniques. In this study, we use magnetic iron oxide nanoparticles for dye removal from
synthetic and real textile wastewater for the rst time. The effects of different experimental parameters on dye removal, such as contact time, initial
concentration, solution pH, and coexisting ions, were investigated. Computational modelling of the interaction of different dye molecules with
different surfaces of g-Fe2O3 nanoparticles is performed to obtain more mechanistic insights on the adsorption behaviour. The results showed that
dye adsorption was fast, as external adsorption was dominated. The adsorption equilibrium data t very closely to the Langmuir adsorption isotherm
model, conrming monolayer adsorption, which is supported by the adsorption computational calculations. The adsorption was spontaneous,
endothermic, and physical in nature.

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63

Journal

* Author to whom correspondence may be addressed.


E-mail address: nassar@ucalgary.ca
Can. J. Chem. Eng. 9999:111, 2015
2015 Canadian Society for Chemical Engineering
DOI 10.1002/cjce.22315
Published online in Wiley Online Library
(wileyonlinelibrary.com).

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

carried out computational modelling to provide insights into the


interaction of the model dye molecules with different maghemite
surfaces. The following factors were investigated to nd the
optimum conditions for enhancing treatment efciency:
solution pH, contact time, initial concentration, coexisting
contaminants, and temperature. This study provides the potential
application of magnetic assisted-nanoparticle technology for dye
removal from textile wastewater efuent.

Table 1. Measured properties of the real wastewater sample considered


in this study
Type of
test

pH

Temperature
(C8)

COD
(ppm)

TOC
(ppm)

Cr3
(ppm)

Value

6.8

25

2582

971

0.0511

those reported in the pdf card 01-073-9835 of the 2005 ICDD


(International Centre for Diffraction Data) database included in
the JADE V.7.5.1 program (Materials Data XRD Pattern Processing
Identication & Quantication).
Adsorbate

FS

A real textile wastewater sample obtained from Al-Aqad Textile


Company (Nablus, Palestine) was used as a source of real
contaminants. The sample was light blue in colour and free of
suspended solid particles. Once the sample reached our lab it was
ltered using a Whatman lter paper to remove any traces of
suspended solids, then the sample pH, total organic carbon (TOC),
chemical oxygen demand (COD), and chromium (Cr3) concentrations were determined (Table 1). The dye structure and
chemistry in the real textile wastewater sample are unknown,
and the disclosure of any identication or structural analysis was
strictly prohibited by the supplier. However, as indicated by the
supplier, the dye present in the real textile wastewater sample is of
an ionic type. Hence, to understand the adsorption mechanism,
two different model dyes, namely crystal violet (CV, C25N3H30Cl,
MW 407.98 g  mol1, lmax 590 nm) and bromocresol green
(BCG, C21H14Br4O5S, MW 698.01 g  mol1, lmax 617 nm),
were obtained from Sigma-Aldrich and used as adsorbates
in preparing synthetic textile wastewater samples. These two
simple dyes have different congurations and contain some of
the functional groups present in dyes that have been used in the
textile industry.[34] Therefore, the model dyes provide insights into
understanding the nature of real textile wastewater and adsorptive
behaviour by the adsorption batch experiment and computational
modelling. The two dyes were used as received without any

EC

TE

Magnetic nanoparticles (g-Fe2O3) purchased from (Alfa Aesar,


ON, Canada) were used as an adsorbent. The nanoparticles were
used as received without further purication. The nanoparticles
have a size of 10  2 nm and surface area of 93 m2/g. Structure and
particle size were determined using an X-ray Ultima III MultiPurpose Diffraction System from Rigaku Corp., (The Woodlands,
TX) with Cu Ka radiation operating at 40 kV and 44 mA with a u-2u
goniometer. Surface area and pore size distribution were
measured using a Tristar II 3020 from Micrometrics (Norcross,
GA) by performing N2-Physisorption at 77 K. Surface area was
calculated using the Brunauer-Emmett-Teller (BET) equation.
External surface area was also measured with the t-plot method,
and there was no signicant difference between the surface areas
obtained by BET and t-plot methods, indicating that the selected
nanoparticles are non-porous. An estimation of the particle size
(assuming spherical particles) was accomplished using the BET
measured specic surface area and the derived equation d 6000/
(SA  roxide);[33] where d is the particle size in nm, SA is the
experimentally-measured specic surface area (93 m2/g), and
roxide is the density of the maghemite oxide (5.49 g/cm3),
producing a value of 11.7 nm which agrees very closely with
the XRD-obtained value.
Figure 1 shows the X-ray diffraction pattern of the selected
nanoparticles. The identication of the pattern conrms the
material to be maghemite, as reported by the manufacturer.
The structure was identied by comparing the XRD signals with

PR

Adsorbent

EXPERIMENTAL

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63

Figure 1. X-ray diffraction pattern of the selected maghemite nanoparticles. Reference data g-Fe2O3 are from Materials Data XRD Pattern Processing
Identication & Quantication.

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

VOLUME 9999, 2015

EC

FS

TE

A batch adsorption method was employed by exposing about


0.10 g of dry mass of g-Fe2O3 nanoparticles to a 10-mL aqueous
solution containing a certain concentration of a specied dye (CV
and BCG). Then, the mixture was shaken vigorously by hand for
a certain time at a pre-determined temperature and time. For the
kinetic study, the aqueous solution containing a 100 mg/L
concentration of each dye and 2582 mg/L COD for the case of
real textile wastewater at 293 K and a specied solution pH (CV:
10.94, BCG: 2.39, RWW: 6.8) was shaken for a specied time
interval. To determine the adsorption time required for equilibrium, samples were selected at predetermined time intervals
and analysed for dye and contaminant concentration. For
pH-dependent studies, adsorption pH experiments were conducted at 293 K and 24 h. 1 mol/L concentration of HNO3 or NaOH
was used to adjust the solution pH in a range of 2.511. For the
isotherm studies, 0.1 g of g-Fe2O3 nanoparticles was added to
10 mL aqueous solution with model dye concentrations ranging
from 0400 mg/L, at different temperatures between 293353 K,
for 24 h. For the case of the real wastewater (RWW) sample, the
isotherm studies were conducted by exposing different masses
of nanoparticles ranging from 0.0250.2 g to a solution of
2582 mg/L COD at the same temperature range and time interval
of the model dyes. For all experiments, after adsorption,
nanoparticles containing the adsorbed dye and other co-existing
ions were separated by a simple magnet, and the supernatant was
decanted and transferred for further analysis. The dye concentration in the synthetic wastewater was measured by UV-vis
spectrophotometry (UV-VIS-NIR-3101PC, Shimadzu). For the
case of RWW samples, the chemical oxygen demand (COD)
measurements were performed instead, following the USEPA
method (USEPA, 2013) and using a Digital Reactor Block DRB
2000 (HACH, Germany). The COD sample was prepared by
adding 1 mL deionized water and 1 mL supernatant (wastewater
sample) to a standard solution from HACH containing 0.65 g/g of
H2SO4, HgSO4, and K2CrO7. Then the solution was transferred to
the DRB instrument for 2 h at 150 8C for COD monitoring. Deionized
water was used as a blank in this test. The adsorbed amount of dye
or COD was determined by mass balance.[20,21] An atomic
absorption instrument iCE 3400 (Scientic Thermo, U.S) was
used for determining Cr3 concentration (a common co-existing
contaminant in textile efuent) before and after adsorption in the
RWW sample. A standard solution containing Cr3 obtained from
Riedel de Hain (Germany) was used for AA calibration. It is worth
noting here that all experiments were performed in duplicate to
conrm reproducibility.

PR

Dye Adsorption onto Nanoparticles

V14[40] and the 3-D molecules optimized with Forcite appear in


Figure 2. As can be seen in Figures 2b and 2d, the two organic
molecules show a 3-D organization that resembles fan blades
and differ from the at conguration on the structural drawings
(Figures 2a and 2c).
The structure of g-Fe2O3 (maghemite) is closely related to the
structure of magnetite (Fe3O4) which possesses the inverse spinel
structure but it differs from magnetite by the presence of iron
vacancies.[35] Vacancy ordering gives rise to different crystal
symmetries, and several possible models for the distribution of
vacancies have been suggested. For this study, we selected the
model with ordered vacancies. Starting with the experimental
structural data reported by Jrgensen, et al.[35] for the tetragonal
maghemite with ordered vacancies, we built the surfaces (001),
(010), (100), and (113). Each surface was created with areas of
approximately 9 nm2 (3 nm  3 nm) to ensure that the selected
organic molecules, CV and BCG, do not interact with their images
in adjacent cells. The depth of the surfaces was set to
approximately 26 A in order to ensure that it was greater than
the non-bond cut-off used in the calculation. The vacuum
thickness was set to 5 nm so that the non-bond calculation for
the CV and BCG molecules did not interact with the periodic image
of the bottom layer of atoms in the surface. The Quality in the
Adsorption Locator calculation was set to Fine, the forceeld
selected was COMPASS, the top layer atoms in each surface were
selected for the interaction with CV and BCG, and the maximum
adsorption distance value was 1.5 nm with a xed energy window
of 418.68 kJ/mol for sampling congurations which differs from
the lowest conguration by the maximum amount. Because of the
large number of possible congurations generated for each surface
arrangement, only the lowest and highest conguration of the two
organic molecules on each surface will be presented and
discussed. To gain further insights into the interaction of CV
and BCG with maghemite nanoparticles, a 10-nm spherical
nanoparticle of maghemite was built and CV and BCG molecules
were allowed to interact with its surface. Two tests were carried
out, one with a single molecule and the second with 160
molecules. Again, the Quality in the Adsorption Locator calculation was set to Fine, the forceeld selected was COMPASS, the
top layer atoms in the spherical nanoparticle surface were selected
for the interaction with CV and BCG, and the maximum adsorption
distance value was 1.5 nm with a xed energy window of
418.68 kJ/mol for sampling congurations.

further purication. Distilled water was used for preparing the


synthetic textile wastewater.

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63

Computational modelling of the interaction of CV and BCG with


maghemite surfaces
Understanding adsorption phenomena is of key importance, and
thus we carried out computational modelling to derive insights
into the interaction of CV and BCG molecules with different
maghemite surfaces. The computational study was done using
Forcite and Adsorption locator within Accelrys Materials Studio
software V7.0.[33,3539] In order to investigate the adsorption of
CV and BCG on maghemite surfaces, the molecules were
structured and optimized with Forcite before adsorbing them
on the selected maghemite surfaces. The quality of the Geometry
Optimization in Forcite was set to Fine and the Forceeld to
COMPASS. The chemical structures drawn with ChemDraw

VOLUME 9999, 2015

RESULTS AND DISCUSSION


Effect of Contact Time (Kinetics Studies)
Figures 3a3b show the dye adsorption experimental data together
with the kinetic model tting for synthetic and real textile
wastewater, respectively, as a function of contact time. As seen,
in all cases, the dye adsorption was reasonably fast: adsorption
equilibrium was reached within 50 min for the model dyes and
< 125 min for the real textile wastewater. This is not surprising, as
the considered maghemite nanoparticles are non-porous and hence
only external adsorption occurs. In contrast, for porous adsorbents
(like activated carbon), adsorption equilibrium time could take
days due to porous diffusion.[13] These ndings are in very close
agreement with those reported by Nassar and coworkers on
the removal of different dyes and metal ions from wastewater using
different metal oxide nanoparticles.[1921,41] To further investigate
the kinetic mechanisms that control the adsorption process, the
experimental data were tted to the Lagergren pseudo-rst-order[42]

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

FS
O
O
PR
D
TE
EC

Figure 2. Chemical structures of (a) crystal violet and (c) bromocresol green; CPK representation of the optimized (b) crystal violet molecule (top and side
views, respectively) and (d) bromocresol green (top and side views, respectively). Grey atoms represent carbon, blue atoms represent nitrogen, white atoms
represent hydrogen, yellow atoms represent sulphur, green atoms represent bromine, and pale green atoms represents chlorine.

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63

Figure 3. Effect of contact time on the adsorptive removal of dye from (a) synthetic wastewater and (b) real textile wastewater. Experimental conditions are
T 293 K, nanoadsorbent dose, 10 g/L; shaking rate, 200 rpm. The symbols are experimental data, the dashed lines are from the pseudo-rst-order model
(Equation (2)), and the solid lines are from the pseudo-second-order model (Equation (3)).

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

VOLUME 9999, 2015

Table 2. Estimated parameters for the pseudo-rst-order and pseudo-second-order models


Pseudo-first-order
Qe (mmol/g)

K1 (min1)

x2

Qe (mmol/g)

K2 (g.mmol1 min1)

10.94
2.39
7.0

0.05
0.03
90.73 (mg/g)

0.28
0.18
0.03

0.00012
0.00027
7.55

0.052
0.026
98.55 (mg/g)

0.35
0.31
0.04

and pseudo-second-order[43] models presented in Equations (1) and


(2), respectively.
dQt
K1 Qe  Qt
dt

dQt
K2 Qe  Qt 2
dt

x2
0.0007
0.0008
4.53

of pollutant at the nanoparticle surface by complexation.[32] It is


worth noting here that the estimated theoretical values (by the
kinetic model) of Qe were in very close agreement to the ones
obtained experimentally.
Effect of Solution pH

FS

EC

TE

where Qe and Qt are the adsorbed amount of pollutant onto the


nanoparticles (mmol/g) at equilibrium and at any time, t (min),
respectively; the equilibrium rate constants K1 and K2 are of
rst-order and second-order adsorption, respectively. The values
of the obtained model parameters and their corresponding
x-values obtained using the Polymath package are presented in
Table 2.
As shown in Figures 3a3b, and on the basis of the x-values
presented in Table 2, both kinetic models t closely to the
experimental data, with the pseudo-rst-order model being the
best t for the model dyes, and the pseudo-second-order model
being the best t for the real textile wastewater sample. This
suggests that the adsorption of the model dye is faster than the real
dye present in the textile wastewater. Generally, the adsorbate
molecule can be transferred from the bulk solution phase to the
adsorbent surface in several steps; including external diffusion,
pore diffusion, surface diffusion, and adsorption on the surface.[13]
Therefore, for the case of non-porous nanoparticles and owing to
an effective degree of mixing and extent dispensability, adsorption
seems to only be affected by electrostatic adsorption (as will be
discussed further in the Effect of Solution pH section).[41] Since
the real textile wastewater sample ts better to the pseudo-secondorder model, it displays two adsorption kinetics: initial rapid
adsorption (presumably due to electrostatic attraction) and slow
adsorption at the later stage which represents a gradual adsorption

The adsorption of wastewater pollutants by maghemite nanoparticles depends signicantly on the electrostatic interactions
between the nanoparticle surface and the functionalized pollutants.
These interactions are inuenced largely by the pH of the solution,
provided that it directly affects the surface charge of the nanoparticles. In this set of experiments, different samples containing a
specied initial concentration (50 mg/L for model dyes in synthetic
wastewater and 2582 mg/L COD of dye in real textile wastewater) at
different initial pH values from 2.511.0 were performed to nd the
optimum adsorption pH. Figures 4a4b show the effect of pH on the
amount of model dyes and real dye from textile wastewater
adsorbed, respectively. Clearly, as seen in Figure 4a for the case of
BCG dye, the maximum amount adsorbed was at pH 3. This
suggests that adsorption is favoured in an acidic environment. On
the other hand, for the case of CV dye, the maximum amount
adsorbed was at pH 10.5, which is favoured in a basic environment. These results are not surprising, as the point of zero charge
(pHpzc) of iron oxide particles is around 7.5,[41,44] and hence ionized
dye adsorption on iron oxide surfaces is likely to be electrostatic
attraction. Therefore, in a relatively basic solution pH > pHpzc, a
signicantly high electrostatic attraction exists between the
negatively-charged surface of iron oxide and the positively-charged
CV dye. On the other hand, as the pH of the solution decreases, the
number of positively-charged sites increases and the number of
negatively-charged sites decreases. This led to a decrease in CV
removal due to charge repulsions. The adsorption process is the
reverse for the BCG dye. It is worth noting here that other
interactions between dye molecules and nanoparticle surfaces

PR

CV
BCG
RWW

pH

Dye

Pseudo-second-order

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63

Figure 4. Effect of pH on adsorptive removal of dye from (a) synthetic wastewater at initial concentration of 50 mg/L; (b) real textile wastewater at initial
COD of 2582.9 mg/L. Experimental conditions are T 293 K, nanoadsorbent dose, 10 g/L; shaking rate, 200 rpm, time 24 h.

VOLUME 9999, 2015

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

Figure 5. Effect of temperature on dye adsorptive removal by g-Fe2O3 nanoparticles. (a) CV dye, (b) BCG dye. Experimental conditions are nanoadsorbent
dose, 10 g/L; shaking rate, 200 rpm, contact time 24 h. The symbols are experimental data, and the solid lines are from the Langmuir model
(Equation (4)).

Effect of Temperature

FS

PR

EC

TE

In this set of experiments, the effect of temperature on adsorptive


dye removal was studied at solution temperatures of 293, 328, and
353 K, and initial concentrations of dyes in synthetic wastewater
were varied from 00.8 mmol/L. For the case of real textile
wastewater, the initial COD concentration was 2582.9 mg/L and
the mass of nanoparticles varied from 0.020.25 g. Figures 5a5b
show that CV and BCG adsorption at the preselected temperatures
increased sharply at a low equilibrium concentration and started
to level off as equilibrium concentration increased. This suggests
that g-Fe2O3 nanoparticles have good adsorption afnity toward
CV and BCG at different temperatures. In addition, increasing the
solution temperature favoured adsorption. This could be attributed to an increase in the mobility of dye molecules and a
subsequent increase in the number of molecules that could
interact with the active adsorption sites on the nanoparticle
surface.[20] Further, the increase of adsorption with temperature
indicates that the adsorption process is endothermic in nature.
Similar observations, but at a slower rate, can be seen for dye
adsorption in real textile wastewater samples (Figure 6). The
slower kinetic rate for the case of real textile dyes agrees closely
with the results obtained in the kinetic study, where the real textile
dye showed second-order kinetic behaviour. In addition, for all the
dyes tested, the experimental adsorption isotherms were tted to
the Langmuir and Freundlich models, represented by Equations (3) and (4), respectively.

where Qe is the amount of dye adsorbed onto nanoparticles


(mmol/g), Ce is the equilibrium concentration of dye in the
supernatant (mmol/L), KL is the equilibrium Langmuir adsorption
constant related to the afnity of binding sites (L/mol), Qmax is
the maximum adsorption capacity for complete monolayer coverage (mmol/g), and KF and 1/n are Freundlich constants which are
related to the adsorbed amount and adsorption afnity, respectively. Both Langmuir and Freundlich model parameters were
estimated by minimizing the sum of squared differences between
the experimental and predicted values using the Solver feature
in
2
eModel
Excel. The non-linear x-square analysis, x2 S QeQQ
,
was
e Model
used to evaluate the closeness of tting results, where Qe and QeModel
are the adsorbed amount of dye obtained experimentally and from
modelling, respectively. The close tting to the experimental data
was indicated by the low x-values.
The estimated values of model parameters are listed in Tables 34.
Clearly, for all the dyes, the Langmuir model presented a closer t to
the experimental data. This suggests that the selected nanoparticles
portray a homogeneous surface where monolayer adsorption
would occur. Clearly, in all cases, the KL and Qmax values increased
with the temperature, which conrms that the adsorption process is
endothermic and favoured with temperature increases.

should not be overlooked, as it will be further discussed in the


Computational Modelling section. However, for the case of the real
textile wastewater sample this apparently was not the case, as the
adsorptive dye removal was not signicantly affected by
solution pH (Figure 4b). This suggests that the dye present in the
real textile wastewater sample might be multifunctional, and
accordingly it could be adsorbed in either basic or acidic environments. The results are interesting for practical application as dye
removal from real textile wastewater efuent could be removed by
magnetic nanoparticles at neutral condition without the need of pH
adjustment.

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63

Qe

Qmax KL Ce
1 KL Ce

Qe KF Ce1=n

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

3
4

Figure 6. Effect of temperature on dye adsorptive removal by g-Fe2O3


nanoparticles from real textile wastewater. Experimental conditions are
initial COD concentration 2582.9 mg/L; shaking rate, 200 rpm, contact
time 24 h. The symbols are experimental data, and the solid lines are
from the Langmuir model (Equation (4)).

VOLUME 9999, 2015

Table 3. Estimated Langmuir parameters at different temperatures. KL (L/mmol), Qmax (mmol/g) for model dyes and KL (L/mg) and Qmax (mg/g) for dye
in real textile wastewater
Temperature (K)
293

328

353

Dyes

pH

KL

Qmax

x2

KL

Qmax

x2

KL

Qmax

x2

CV
BCG
RWW

10.94
2.39
6.8

20.99
9.32
0.00002

0.49
0.19
363.6

0.050
0.005
1.43

20.04
10.63
0.00015

0.55
0.20
413.2

0.04
0.01
1.71

21.26
12.03
0.0002

0.61
0.21
425.2

0.04
0.01
2.46

Table 4. Estimated Freundlich parameters at different temperatures. 1/n (unit less), KF [(mmol/g)(L/mmol)1/n] for model dyes and KF [(mg/g)
(L/mg)1/n] for dye in real textile wastewater
Temperature (K)
328 K

353 K

pH

KF

1/n

x2

KF

1/n

x2

CV
BCG
RWW

10.94
2.39
6.8

0.9
0.77
0.17

0.5
0.84
0.98

0.2
0.19
1.43

0.89
0.84
0.52

0.49
0.84
0.89

0.17
0.22
1.71

KF

1/n

x2

0.87
0.84
0.35

0.49
0.81
0.85

0.16
0.20
2.46

Dyes

FS

293 K

PR

TE

To better understand the effect of temperature on adsorptive dye


removal, thermodynamic studies were employed by estimating
the changes in standard Gibbs free energy (DG8), enthalpy (DH8),
and entropy (DS8) of adsorption. The values of DG8 were estimated
using Equation (5).

of an individual one. As a result, the effect of coexisting pollutants


should be addressed when conducting an adsorption study. For the
case of real textile wastewater it appears that in addition to
removing colour from textile wastewater, the nanoparticle
successfully removed other contaminates, such as Cr3, at a rate
of more than 98 %, as conrmed by the atomic absorption analysis.

Thermodynamic Studies

DGo RT ln K

Computational Modelling

Figures 89 show some congurations that the CV and BCG dye


molecules may obtain after adsorption on different maghemite
surfaces. In Figure 8, it is possible to see that the original fan
blades conguration of CV (Figure 2b) is modied upon
adsorption to the different surfaces of maghemite, where a
attening of the molecule is clearly observed to occur as it interacts
with the different selected surfaces of maghemite. According to the

lnK 

DHo DSo

RT
R

EC

where R is the universal ideal gas constant (R 8.314 J/mol  K), T


is the temperature (K), and K is the adsorption equilibrium
constant. K is expressed as KL  Cs, where KL is the Langmuir
equilibrium constant (L/mmol) and Cs is the solvent molar
concentration (mmol/L), which is estimated from the density and
molecular mass of water.[45]
The values of DH8 and DS8 were estimated using the Van t Hoff
equation[46] (Equation (6)), by plotting ln(K) against 1/T
(Figure 7). DH8 and DS8 could be measured from the slope and
intercept of the best-t line, respectively. The calculated
thermodynamic parameters for the model dyes and real dye in
textile wastewater are presented in Table 5.

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63

As seen in Table 5, for all dyes, the values of DG8 at all


temperatures are negative, which indicates that dye adsorption
onto nanoparticle surfaces is spontaneous and thus thermodynamically favourable. Further, the DH8 value is positive for all
dyes, which implies that dye adsorption is endothermic in nature.
The positive value of DS8 may be attributed to the increase in
randomness at the nanoparticle surfaceliquid interface.[20,21]
Effect of Coexisting Pollutants
As efuent wastewater contains more than one pollutant, the
presence of other pollutants may interfere in the removal efciency

VOLUME 9999, 2015

Figure 7. van't Hoff plot for the endothermic adsorption of CV, BCG and
RWW.

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

Table 5. Values of DG8, DH8 and DS8 for adsorptive removal of CV, BCG and dye in real wastewater samples at different temperatures. DG8:(kJ/mol), DS8:
(J/mol.K), DH8:(kJ/mol), K:(unitless)
Temperature (K)
293
Dye

353

4G8

4S8

4H8

K x103

4G8

K x103

DG8

Kx103

34.03
32.05
6.95

116.14
128.2
144.17

0.05
5.5
35

1164.52
517.32
0.17

35.64
34.01
13.13

1106.68
587.27
0.12

38.09
36.54
15.32

1163.55
658.84
0.19

PR

FS

CV
BCG
RWW

353

EC

TE

Figure 8. CPK images of the adsorption of crystal violet on the surfaces of maghemite (side and top views, respectively). (a) Lowest and highest
conguration on surface (001); (b) lowest and highest conguration on surface (010), and (c) lowest and highest conguration on surface (100); and (d)
lowest and highest conguration on surface (113). Bright blue atoms represent nitrogen, gray atoms represent carbon, white atoms represent hydrogen,
pale green atoms represent chlorine, red atoms represent oxygen and light blue atoms represent iron.

adsorption calculations, CV has a strong interaction with the


maghemite surfaces through nitrogen atoms and aromatic rings,
forcing the at conguration of CV on the surface. The adsorption
energy of CV on the (010) surface is the lowest of the four studied
surfaces, and that on the surface (113) is the highest: a difference
of 1105.31 kJ /mol was obtained between the lowest congurations on surface (010) and on surface (113) indicating that
surface (113) is the least favourable for CV adsorption. Table 6
shows the differences in adsorption energy with respect to the
lowest surface adsorption on surface (010). Differences between

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63

Table 6. Adsorption energy differences for the adsorption of CV and


BCG in the studied surfaces of maghemite with respect to the surface
(010)
CV Adsorption Energy
difference with respect to
(010) surface [kcal/mol]

BCG Adsorption Energy


difference with respect to
(010) surface [kcal/mol]

Surface

Lowest

Highest

Lowest

Highest

(010)
(001)
(100)
(113)

0
22.98
3.52
264.13

92.97
76.96
99.96
332.72

0
7.99
3.90
100.72

99.83
107.91
103.79
200.01

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

the values of adsorption energy of CV for surfaces (010), (100),


and (001) are smaller, predicting that the adsorption of CV on
these maghemite surfaces should follow the Langmuir adsorption
model, which indeed was observed experimentally, and that the
surface (113) must grow to be exposed in larger areas in order to
modify the adsorption model from Langmuir to Freundlich.
Figure 9 illustrates the conguration of the BCG molecule as it is
adsorbed on the different studied maghemite surfaces. For this
case, it is observed that the BCG molecule interacts with the
surface through aromatic rings and hydroxyl groups, and the
sulfur double-bonded with oxygen atom points upward in
the lowest and highest congurations. The congurations are
almost at compared with the original conguration (Figure 2d).
As happened with the CV molecule, the lowest adsorption energy
is found on surface (010) and the highest on surface (113). A
difference of 418.68 kJ/mol is observed between the lowest
adsorption energies on surface (010) and surface (113), as shown
in Table 6. Again, the differences in adsorption energies between
surfaces (010), (100), and (001) are small, indicating that BCG and
CV do not sense substantial differences in the adsorption sites on
these surfaces. Thus, a Langmuir adsorption model is predicted
according to these results. This was observed experimentally
when CV and BCG were adsorbed on maghemite nanoparticles.
Table 7 shows a summary of the congurations that CV and BCG
can obtain on the studied maghemite surfaces.

VOLUME 9999, 2015

FS

Figure 9. CPK images of the adsorption of bromocresol green on the surfaces of maghemite (side and top views, respectively). (a) Lowest and highest
conguration on surface (001); (b) lowest and highest conguration on surface (010), and (c) lowest and highest conguration on surface (100); and (d)
lowest and highest conguration on surface (113). Bright green atoms represent bromine, yellow atoms represent sulfur, gray atoms represent carbon,
white atoms represent hydrogen, red atoms represent oxygen and light blue atoms represent iron.

Surface

Highest

Flat
Flat
Flat
Flat

Flat-slightly tilted
Flat-slightly tilted
Flat-slightly tilted
Flat

Lowest
Almost
Almost
Almost
Almost

flat
flat
flat
flat

Highest
Almost
Almost
Almost
Almost

flat
flat
flat
flat

TE

(010)
(001)
(100)
(113)

Lowest

BCG (Configuration)

PR

CV (Configuration)

on a 10 nm spherical maghemite nanoparticle (closest to the


actual size obtained experimentally for the real maghemite
sample used). Figures 1011 illustrate the adsorption of
one molecule of CV or BCG and the adsorption of 160 molecules
of CV or BCG on the 10 nm spherical maghemite nanoparticle,
respectively. In Figure 11a (expanded insets) it is possible to see
that in the rounded surface of the 10 nm spherical maghemite
nanoparticle, the lowest conguration of CV is no longer the at
one observed in the perfectly at surfaces of maghemite, but a
tilted one (more vertical) anchored by the interaction of two
nitrogen atoms that are not positively charged (this conguration
was observed in the at surfaces but the adsorption energy was
high, and thus not discussed in the previous section). The
positively-charged nitrogen is pointing upwards together with the
chloride atom, in the absence of water. In Figure 10b it is observed
that as more CV molecules are added (in this case 160), the tilting

Table 7. Summary of the different congurations of CV and BCG in the


different maghemite surfaces

EC

The previous CV and BCG adsorption calculations were carried


out on perfectly at maghemite surfaces. Thus, to get a glimpse of
what may actually be happening with a real nanoparticle,
calculations of the adsorption of CV and BCG were carried out

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63

Figure 10. CPK images of the adsorption of crystal violet on the surfaces of a 10 nm maghemite nanoparticle. (a) side and top views of the adsorption of one
molecule of CV on the 10 nm maghemite nanoparticle; and (b) adsorption of 160 CV molecules on the 10 nm maghemite nanoparticle. Bright blue atoms
represent nitrogen, gray atoms represent carbon, white atoms represent hydrogen, pale green atoms represent chlorine, red atoms represent oxygen and
light blue atoms represent iron.

VOLUME 9999, 2015

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

PR

on the initial concentration of soluble contaminants. The presence


of nanoparticles in the bed shall not impact the bed permeability nor
ux, as the nanoparticles will be integrated with the sand grains or
other media to support and hold the nanoparticles, and will be used
in small quantities (less than 0.1 g of nanoparticles per gram of bed
material).[30] Another advantage of this packed bed process is that
after adsorption, the nanoparticles can be employed as catalysts,
whereby the adsorbed organic contaminants can be gasied into
synthetic gas with the aid of steam (catalytic steam gasication
process).[47] This will also help in the regeneration of nanoparticles
for another cycle or dumping them safely without any negative
impact on the environment, as sand- and iron-based nanoparticles
are naturally occurring and environmentally-friendly. We believe
that incorporating nanoparticle technology in conventional wastewater treatment is an important step forward in making conventional wastewater treatment plants more accessible to different
types of industrial efuents and not just municipal wastewater, as is
the case in Palestine. By incorporating nanoparticle technology
with conventional wastewater treatment processes, this technology
becomes overwhelmingly cost-effective and appealing to the
average wastewater producer.

EC

TE

of the CV molecules is more pronounced (more vertical) than


when one lone molecule of CV is on the nanoparticle surface. This
arrangement indicates that more CV molecules can be adsorbed
on the surface than would be estimated by the at conguration of
the molecule observed for the at surfaces studied, pointing out
the importance of surface morphology for CV adsorption.
Figure 11 shows the lowest conguration of BCG adsorption on
the 10 nm maghemite nanoparticle. For this case, the conguration
of BCG on the rounded nanoparticle surface (Figure 11a) is similar
to that observed for the at surfaces in which aromatic rings and
hydroxyl groups are the anchoring sites toward the surface. When
more BCG molecules are added (160 in this case), it is possible to see
that some of the BCG molecules can tilt slightly on the surface of the
spherical nanoparticle (Figure 11b), but not all of them are tilted as
in the case of CV molecules. This indicates that the conguration of
BCG molecules on the surface remains similar to when only one is
present, which should predict a difference in adsorption capacity
with respect to CV. This supports the experimental ndings on
kinetic and isotherm studies (Figures 3 and 5). Both gures show
that CV uptake is greater than BCG uptake, which may be related to
the insights gained in the adsorption calculations carried out for the
spherical maghemite nanoparticles.

FS

Figure 11. CPK images of the adsorption of bromocresol green on the surfaces of a 10 nm maghemite nanoparticle. (a) side and top views of the adsorption
of one molecule of BCG on the 10 nm maghemite nanoparticle; and (b) adsorption of 160 BCG molecules on the 10 nm maghemite nanoparticle. Bright
green atoms represent bromine, yellow atoms represent sulfur, gray atoms represent carbon, white atoms represent hydrogen, red atoms represent oxygen
and light blue atoms represent iron.

TECHNOLOGY

SCALING-UP AND COST-EFFECTIVENESS OF NANOPARTICLE

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63

Our study on the application of nanoparticles falls under the eld of


in situ remediation rather than packed bed processes. As
mentioned, due to high medium mobility with high dispersion
ability, nanoparticles lead to effective transport for in situ
applications. Therefore, the nanoparticles can be employed in
situ, within the contaminated zone. Nevertheless, integrating
nanoparticle technology with conventional wastewater treatment
processes is possible.[30] Nanoparticle technology can be integrated
with the sedimentation process, were it can work as both a
coagulant and an adsorbent. Further, nanoparticle technology can
be integrated with bed ltration technology, where the ltration bed
will trap suspended contaminants and the nanoparticles will adsorb
soluble ones.[30] The ratio of nanoparticles to bed media will depend

10

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

CONCLUSIONS
Iron oxide nanoparticles have shown potential for environmental
applications because of their unique properties, such as specic
functionality and large specic surface area per unit mass. The
results from the present study indicate that g-Fe2O3 nanoparticles
could be employed successfully for decolourization of textile
wastewater efuent. Dye adsorption was fast and equilibrium was
achieved in less than 125 min for dye in real textile wastewater.
Further, for dye and contaminants in real textile wastewater,
adsorption was favoured at higher temperatures and was not
signicantly affected by solution pH, indicating dye multifunctionality. It was found that the equilibrium data closely t
the Langmuir model, suggesting monolayer adsorption, which
was backed up by the adsorption calculations. The adsorption
results of the thermodynamics studies showed the endothermic
nature of the adsorption process, the spontaneity, and the

VOLUME 9999, 2015

ACKNOWLEDGEMENTS

EC

PR

TE

[1] B. Shathika Sulthana Begum, R. Gandhimathi, S. T. Ramesh,


P. V. Nidheesh, J. Mater. Cycles Waste Manage 2013, 15, 564.
[2] S. Khorramfar, N. M. Mahmoodi, M. Arami, K. Gharanjig,
Color. Technol. 2010, 126, 261.
[3] C. Bradu, L. Frunza, N. Mihalche, S. M. Avramescu,
M. NeaSt a, I. Udrea, Appl. Catal., B. 2010, 96, 548.
[4] Z. Ai, J. Li, L. Zhang, S. Lee, Ultrason. Sonochem 2010, 17,
370.
[5] A. Afkhami, R. Moosavi, J. Hazard Mater. 2010, 174, 398.
[6] S. Yang, H. He, D. Wu, D. Chen, Y. Ma, X. Li, J. Zhu, P. Yuan,
Ind. Eng. Chem. Res. 2009, 48, 9915.
[7] G. Moussavi, M. Mahmoudi, J. Hazard Mater. 2009, 168, 806.
[8] E. Eren, J. Hazard Mater. 2009, 162, 1355.
[9] H.-Y. Zhu, R. Jiang, L. Xiao, W. Li, J. Hazard Mater. 2010,
179, 251.
[10] R. Salehi, M. Arami, N. M. Mahmoodi, H. Bahrami,
S. Khorramfar, Colloids Surf., B 2010, 80, 86.
[11] F. Emami, A. R. Tehrani-Bagha, K. Gharanjig, F. M. Menger,
Desalination 2010, 257, 124.
[12] M. S. Nawaz, M. Ahsan, Alexandria Eng. J. 2014, 53, 717.
[13] Metcalf & Eddy, Wastewater engineering treatment and
reuse, 4th edition, McGraw Hill, New York 2003.
[14] W. Lei, D. Portehault, D. Liu, S. Qin, Y. Chen, Nat. Commun.
2013, 4, 1777.
[15] J. Hu, G. Chen, I. M. C. Lo, Water Res. 2005, 39, 4528.
[16] J. Hu, G. Chen, I. M. C. Lo, J. Environ. Eng. 2006, 132, 709.
[17] J. Hu, Lo, G. Chen, Langmuir 2005, 21, 11173.
[18] J. Hu, I. M. C. Lo, G. Chen, Sep. Purif. Technol. 2007, 56, 249.
[19] N. N. Nassar, Separ. Sci Technol. 2010, 45, 1092.
[20] N. N. Nassar, J. Hazard Mater. 2010, 184, 538.
[21] N. N. Nassar, Can. J. Chem. Eng. 2012, 90, 1231.
[22] N. N. Nassar, A. Hassan, P. Pereira-Almao, Energy Fuels
2011, 25, 1017.

REFERENCES

The authors thank Dr. A. Abu Obaid from the Department of


Chemistry for providing the model dyes, Mr. N. Dwikat for his
assistance in the UV-vis and AA analyses and the Department
of Chemical Engineering for providing the COD kits, and
Mrs. Hamees Tbaileh and Mr. Yusef Ratrout for their help in
COD measurements. Special thanks to Al-Aqad Textile Company
in Nablus, Palestine for providing the textile wastewater sample.
The authors would like to thank the Engineer Zuhair Hijjawi
Award 2014. Dr. Nassar thanks the Palestinian American Research
Center for the 2014/2015 fellowship. The authors are also grateful
to the Natural Sciences and Engineering Research Council of
Canada (NSERC), Nexen-CNOOC Ltd, and Alberta InnovatesEnergy and Environment Solutions (AIEES) for the for the
nancial support provided through the NSERC/NEXEN/AIEES
Industrial Research Chair in Catalysis for Bitumen Upgrading.

[23] J. M. Perez, Nat. Nanotechnol. 2007, 2, 535.


[24] G. A. Waychunas, C. S. Kim, J. F. Baneld, J. Nanopart. Res.
2005, 7, 409.
[25] N. Savage, M. S. Diallo, J. Nanopart. Res 2005, 7, 331.
[26] G. Schmid, Nanoparticles: From Theory to Application,
Wiley-VCH, Weinheim 2004.
[27] M. C. Roco, Curr. Opin. Biotechnol. 2003, 14, 337.
[28] C. M. Niemeyer, Angew. Chem. Int. Ed. 2001, 40, 4128.
[29] J. F. Baneld, H. Zhang, Nanoparticles in the environment,
in Nanoparticles and the Environment, P. H. Ribbe,
J. J. Rossi, Eds., The Mineralogical Society of America,
Washington, D. C. 2001, pp. 1200.
[30] N. N. Nassar, L. A. Arar, N. N. Marei, M. M. Abu Ghanim,
M. S. Dwekat, S. H. Sawalha, Environ. Nanotech. Monit.
Manage. 2014, 12, 14.
[31] P. G. Tratnyek, R. L. Johnson, Nano Today 2006, 1, 44.
[32] N. N. Nassar, Iron Oxide Nanoadsorbents for Removal of
Various Pollutants from Wastewater: An Overview, in
Application of Adsorbents for Water Pollution Control,
A. Bhatnagar, Ed., Bentham Science Publishers, Sharjah,
UAE 2012, p. 81118.
[33] N. N. Nassar, A. Hassan, G. Vitale, Appl. Catal. A 2014, 484,
161.
[34] A. Ghaly, R. Ananthashankar, M. Alhattab, V. Ramakrishnan, J. Chem. Eng Process. Tec. 2014, 5, 182.
[35] J.-E. Jrgensen, L. Mosegaard, L. E. Thomsen, T. R. Jensen,
J. C. Hanson, J Solid State. Chem. 2007, 180, 180.
[36] S. Kirkpatrick, C. D. Gelatt Jr, M. P. Vecchi, Optimization by
Simulated Annealing, in Readings in Computer Vision,
M. A. Fischler, O. Firschein, Eds., Morgan Kaufmann,
San Francisco 1987, p. 606615.
[37] J. Barriga, B. Coto, B. Fernandez, Tribol Int. 2007, 40, 960.
[38] O. Ermer, Calculation of molecular properties using force
elds: Applications in organic chemistry, in Bonding forces,
M. Simonetta, A. Gavezzotti, K. D. Warren, O. Ermer,
Springer Berlin Heidelberg, Germany 1976, p. 161211.
[39] Accelrys, Inc., Materials Studio Modeling and Simulation
Software Version 7.0, 2014.
[40] CambridgeSoft Corporationa, subsidiary of PerkinElmer, Inc,
ChemDraw V14 Structural Drawing Software, 2014.
[41] N. N. Nassar, A. Ringsred, Environ. Eng. Sci. 2012, 29, 790.
[42] Y. S. Ho, Scientometrics 2004, 59, 171.
[43] Y. S. Ho, G. McKay, Process Saf. Environ. 1998, 76, 332.
[44] L. S. Balistrieri, J. W. Murray, Am. J Sci. 1981, 281, 788.
[45] A. Rudrake, K. Karan, J. H. Horton, J. Colloid Interface Sci
2009, 332, 22.
[46] J. M. Smith, H. C. VanNess, M. M. Abbott, Introduction to
chemical engineering thermodynamics McGraw Hill, New
York 2005.
[47] N. N. Nassar, A. Hassan, P. Pereira-Almao, Energy Fuels
2011, 25, 1566.

FS

physisorption of the process. As a conclusion, this study conrmed


that g-Fe2O3 nanoparticles could be employed as an efcient
adsorbents for industrial wastewater treatment where classical
adsorbents could be ineffective or costly.

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63

VOLUME 9999, 2015

Manuscript received November 18, 2014; revised manuscript


received January 19, 2015; accepted for publication February 23,
2015.

THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING

11

AUTHOR QUERY FORM


JOURNAL: THE CANADIAN JOURNAL OF CHEMICAL ENGINEERING
Article: cjce22315
Dear Author,
During the copyediting of your paper, the following queries arose. Please respond to these by annotating your
proofs with the necessary changes/additions using the E-annotation guidelines attached after the last page of this
article.
We recommend that you provide additional clarication of answers to queries by entering your answers on the
query sheet, in addition to the text mark-up.
Query

Remark

Q1

Please conrm that given names (red) and surnames/family names (green)
have been identied correctly.

EC

TE

PR

FS

Query No.

FS

Color figures were included with the final manuscript files that we received for your article. Because of the
high cost of color printing, we can only print figures in color if authors cover the expense. The charge for
printing figures in color is $600 per figure.

Please indicate if you would like your figures to be printed in color or black and white. Color images will be
reproduced online in Wiley Online Library at no charge, whether or not you opt for color printing.

PR

Failure to return this form will result in the publication of your figures in black and white.
VOLUME

JOURNAL

NO. OF
COLOR PAGES

AUTHOR(S)

EC

MS. NO.

TE

TITLE OF MANUSCRIPT

Address

Institution

BILL TO:
Name

Please print my figures in color

Please print my figures in black and white

$
Purchase
Order No.
Phone

Fax
E-mail

ISSUE

Вам также может понравиться