You are on page 1of 35

Journal Pre-proof

Adsorption performance of UiO-66 towards organic dyes: Effect


of activation conditions

Danial Vaghar Mousavi, Salman Ahmadipouya, Atefeh


Shokrgozar, Hossein Molavi, Mashallah Rezakazemi, Farhad
Ahmadijokani, Mohammad Arjmand

PII: S0167-7322(20)34528-1
DOI: https://doi.org/10.1016/j.molliq.2020.114487
Reference: MOLLIQ 114487

To appear in: Journal of Molecular Liquids

Received date: 11 July 2020


Revised date: 29 September 2020
Accepted date: 30 September 2020

Please cite this article as: D.V. Mousavi, S. Ahmadipouya, A. Shokrgozar, et al.,
Adsorption performance of UiO-66 towards organic dyes: Effect of activation conditions,
Journal of Molecular Liquids (2018), https://doi.org/10.1016/j.molliq.2020.114487

This is a PDF file of an article that has undergone enhancements after acceptance, such
as the addition of a cover page and metadata, and formatting for readability, but it is
not yet the definitive version of record. This version will undergo additional copyediting,
typesetting and review before it is published in its final form, but we are providing this
version to give early visibility of the article. Please note that, during the production
process, errors may be discovered which could affect the content, and all legal disclaimers
that apply to the journal pertain.

© 2018 Published by Elsevier.


Journal Pre-proof

Adsorption Performance of UiO-66 towards Organic Dyes:


Effect of Activation Conditions

Danial Vaghar Mousavi1, Salman Ahmadipouya1, Atefeh Shokrgozar1, Hossein Molavi1*,


Mashallah Rezakazemi2, Farhad Ahmadijokani3*, Mohammad Arjmand3*
1
Department of Chemical and Petroleum Engineering, Sharif University of Technology,

Tehran, Iran

of
2
Faculty of Chemical and Materials Engineering, Shahrood University of Technology,

ro
Shahrood, Iran
3
School of Engineering, University of British Columbia, Kelowna, BC, V1V 1V7 Canada
-p
Corresponding authors: Farhad Ahmadijokani, Email: farhad.ahmadijokani@ubc.ca,
re
Hossein Molavi, Email: hossein.molavi128@gmail.com and Mohammad Arjmand, Email:
mohammad.arjmand@ubc.ca
lP
na
ur
Jo

1
Journal Pre-proof

Abstract

The Zr-based metal-organic framework (MOF, UiO-66) was synthesized solvothermally.

Then, the synthesized UiO-66 was activated using different solvents (acetone, chloroform,

and ethanol) via two activation methods of centrifugation and Soxhlet extraction over

different periods (1-10 days). The crystalline structure and morphology of the synthesized

UiO-66s were characterized by X-ray diffraction (XRD), nitrogen adsorption-desorption,

of
and field-emission scanning electron microscopy (FESEM) techniques. The adsorption

ro
behaviors of the synthesized UiO-66s were then investigated by selecting anionic methyl

-p
red (MR) and cationic methylene blue (MB) as the model dyes. It was found that a change

in activation solvent and activation method would alter the physical properties of UiO-66,
re
such as surface area, pore-volume, pore diameter, and surface charge. Experimental
lP

adsorption tests demonstrated that UiO-66 activated by ethanol via the Soxhlet extraction
na

method for five days (abbreviated as ES-5) presented higher MR uptake capacity (243

mg/g) than the other synthesized UiO-66s, ascribed to its large surface area and total pore
ur

volume. Furthermore, it was found that the uptake capacity of ES-5 for cationic MB and
Jo

anionic MR dyes increased and decreased, respectively, with raising the solution pH. This

resulted in superior adsorption selectivity of MB over MR at pH 9.1. All in all, the findings

of this work showed that the Soxhlet extraction method holds great promise for the

activation of different MOFs.

Keywords: Metal-organic framework; UiO-66; Activation solvent; Selective adsorption;


Dye removal

2
Journal Pre-proof

1. Introduction

Discharge of colored wastewater into the surface waters is increasing unceasingly, mainly

due to the growth of industrial activities, technological development, and intensively used

organic dyes as traditional colorants in the paper, leather, printing, plastic, and other

industries [1, 2]. Organic dyes and their derivatives are highly toxic and non-biodegradable,

posing severe environmental problems [3, 4]. Because of the mentioned harmful effects, the

of
removal of organic dyes from wastewater has attracted the attention of researchers. Among

ro
different methods, adsorption has received significant recognition for the removal of

-p
organic dyes from wastewater due to its low operation cost, high efficiency, and facile
re
regeneration of exhausted adsorbents [3, 5].
lP

Metal-organic framework (MOF) is an emerging class of hybrid nanoporous materials built

up from metal ions and organic bridging linkers [6, 7]. MOFs have realized a broad
na

spectrum of applications such as gas separation and storage, dye adsorption, heavy metal
ur

adsorption, drug adsorption and delivery, catalysis, sensors, and so on [8-11]. As such,

recently, MOFs have been researched intensely in chemistry and environmental science
Jo

[12, 13]. Moreover, the applications of MOF nanomaterials can be extended due to their

large surface area, highly ordered porosities, tunable surface functionalities, pore size and

shape tenability, and low densities [14, 15].

Over the past decade, numerous types of MOFs have been employed for dye adsorption and

separation. UiO-66 is a member of the MOF family well-known for its excellent thermal

3
Journal Pre-proof

and chemical stability and large surface area; hence, it has been intensively studied for the

adsorption and separation of organic dyes from aqueous solutions [16-19]. In the structure

of UiO-66, the Zr6O4(OH)4 clusters are interconnected by twelve terephthalate linkers to

form an extended face-centered-cubic (fcc) topology. Such a framework results in one of

the largest coordination networks among all MOFs, which is known to have excellent

structural stability under harsh conditions [20, 21].

of
Although the adsorption performance of UiO-66 and its derivatives for the removal of

ro
organic dyes has been widely studied; nevertheless, the activation methods and their effects

-p
on adsorption performance have rarely been explored. Understanding the impact of

physical properties, such as surface area, pore-volume, and surface charge of UiO-66, on
re
the adsorption performance is very important because it is expected that these parameters
lP

can alter the uptake capacity of organic dyes. Usually, activation solvents and activation
na

methods are two of the main parameters that can change these properties of UiO-66.

However, the literature holds a handful number of studies devoted to the activation of UiO-
ur

66 by solvent-exchange using different solvents via the centrifugation method.


Jo

Accordingly, activation solvent with low boiling point and poor interactions with pore

walls of MOFs such as acetone, chloroform, methanol, ethanol, and dichlorocarbon are

suitable solvents for the activation process [22]. Therefore, based on the properties

mentioned above, in the present work, three solvents, including acetone, chloroform, and

ethanol, were used as activation solvent.

Activation of UiO-66 by soaking in different solvents for a long time and then

4
Journal Pre-proof

centrifugation features some restrictions, such as being costly and time-consuming and

offering low material production efficiency. Thus, it is much desired to increase the surface

area and pore volume of UiO-66 nanoparticles by a facile swift activation method[23]. So,

in the current research, the synthesized MOFs were activated via two different conventional

activation methods like centrifugation and Soxhlet extraction methods. Then the adsorption

behaviors of various UiO-66s activated by different activation solvents and methods were

of
studied by choosing methyl red (MR) and methylene blue (MB) as the model dyes. In this

regard, the effects of governing parameters such as agitation time, initial dye concentration,

ro
temperature, and solution pH were investigated. Additionally, the relationships between

-p
physical properties (such as surface area, pore-volume, and surface charge) and uptake
re
capacity were studied in detail.
lP
na

2. Experimental
ur

2.1. Materials
Jo

Dimethylformamide (DMF, Merck), ZrCl4 (Sigma–Aldrich), terephthalic acid (TPA,

Merck), methylene blue (MB, Merck), methyl red (MR, Merck), chloroform (analytical

grade), ethanol (Merck), and acetone (analytical grade) were used as received from the

vendors without any purification.

2.2. Synthesis

A modified and scaling-up synthesis procedure was used to synthesize UiO-66[17]. Briefly,

5
Journal Pre-proof

760 mg ZrCl4, 1060 mg TPA, and 62 mL DMF were mixed at room temperature and

consequently sonicated for 30 min to completely dissolve all solid materials. Then, the

solution was transferred into an autoclave and heated for 24 h at 393 K in a hot air oven.

After being cooled down naturally to room temperature, the synthesized nanomaterials

were collected in two different ways, i.e., filtration or centrifugation, and then activated via

two different techniques. i.e., solvent exchange via centrifugation and Soxhlet extraction.

of
The collection and activation phases are elaborated in the supporting information (Section

S1). Finally, the synthesized nanomaterials were dried under vacuum at 393 K for 18 h.

ro
Table S1 summarizes the activation conditions and the textural properties of the developed

-p
nanomaterials. The synthesized UiO-66s are referred as XY-d, which X refers to the
re
activation solvents (A, C and E, as the first letter, denote acetone, chloroform, and ethanol,
lP

respectively), Y relates to the activation method (S and C, as the second letter, signify

Soxhlet extraction and centrifugation, respectively), and d refers to the time of activation
na

(1-10 day).
ur
Jo

2.3. Analysis

The UiO-66 activated with different methods was characterized by XRD, BET, FESEM,

and zeta potential analysis, where the details are delineated in the supporting information

(Section S2).

6
Journal Pre-proof

3. Results and Discussion

3.1. Materials Characterization

XRD analysis was used to confirm the synthesis of UiO-66 nanoparticles under different

activation methods and various activation solvents, and the obtained XRD patterns are

shown in Figures 1 and S1. The XRD profiles of UiO-66 activated in chloroform and

ethanol via both centrifugation and Soxhlet extraction methods were similar and in good

of
agreement with the patterns reported in the literature [24-26]. This corroborates the

ro
successful synthesis of this MOF featuring the same crystalline structure as the literature.

-p
However, as can be observed in Figure 1a, the XRD patterns of UiO-66 activated by
re
acetone via the Soxhlet extraction method (AS-5) were different than other synthesized

MOFs. Figure 1a reveals several weak and broad peaks at 2θ = 6-12°, 14-20°, and 22-37°,
lP

suggesting an amorphous structure for AS-5. This observation can be related to the collapse
na

of the crystalline structure of UiO-66 after being immersed in the Soxhlet setup for five

days.
ur

In contrast to AS-5, the XRD patterns of UiO-66 activated by acetone via the centrifugation
Jo

method (AC-5) displayed all of the characteristic peaks of UiO-66, comparable to the XRD

patterns of other synthesized UiO-66s. But, the peak intensities of this MOF decreased

significantly as compared to CC-5 and EC-5. Furthermore, comparing Figures 1a and 1b

showed that the peak intensities of UiO-66 activated via the centrifugation method is

remarkably higher than those activated via Soxhlet extraction method. This suggests the

successful activation process and crystalline structures preservation using the centrifuge

7
Journal Pre-proof

method. Therefore, based on the obtained results, it can be said that among the three

activation solvents (acetone, ethanol, and chloroform), ethanol presents the best

performance, as its corresponding UiO-66s show the highest crystallinity.

of
ro
-p
re
lP
na
ur
Jo

8
Journal Pre-proof

of
ro
-p
re
lP
na
ur
Jo

9
Journal Pre-proof

Figure 1. XRD patterns of UiO-66 activated via different solvents (acetone, ethanol and

chloroform) and methods: (a) Soxhlet extraction and (b) Centrifugation. A, C and E, as the

first letter, denote acetone, chloroform, and ethanol, respectively. S and C, as the second

letter, signify Soxhlet extraction and centrifugation, respectively.

of
The impact of activation time on the crystalline structure of UiO-66 activated via these

ro
activation methods was also studied. In this regard, acetone was selected as an activation

solvent because of its strong effect on the crystalline structure of UiO-66, as observed in
-p
Figure 1. As depicted in Figure S1, the UiO-66s activated via centrifugation method could
re
maintain their crystalline structures in acetone from 1 to 10 days. Still, their peak intensities
lP

decreased slightly with the increase of activation time. However, the XRD patterns of UiO-

66 activated via the Soxhlet extraction method for longer periods (5 and 10 days) showed
na

relatively weak and broad peaks, suggesting that acetone-activated MOFs using the Soxhlet
ur

extraction method have low crystallinity and an amorphous structure. These observations
Jo

indicate the preservation and disruption of the crystalline structure of UiO-66 against

acetone in centrifugation and Soxhlet extraction methods, respectively. As reported

previously by Cavka et al.[27], it is worthy to note that UiO-66 showed excellent structural

stability against acetone at room temperature for 24 h, which is in harmony with our results.

Nevertheless, it should be said that the stability of the synthesized UiO-66 against boiling

acetone in the Soxhlet setup is low, and its crystalline structure collapsed after being

immersed in acetone for more than one day.

10
Journal Pre-proof

Moreover, the structural stability of the synthesized UiO-66 against water was investigated

by stirring these MOFs in water for 7 days. Based on the results obtained form XRD

analysis (Figure S2), it can be found that the water stability of all synthesized UiO-66s at

this time is the same and do not show significant crystalline change after being stirred in

water for 7 days.

The nitrogen adsorption-desorption isotherms, pore size distributions of the synthesized

of
MOFs, and the measurement parameters are shown in Figures 2, S3, and S4, and Table S2,

ro
respectively. All of the synthesized MOFs showed I-type isotherms (except for AS-5 and

-p
AS-10), demonstrating the existence of micropores for AS-1, AC-1, AC5, AC-10, CS-5,

ES-5, CC-5, and EC-5. The type IV isotherms for AS-5 and AS-10 verified the generation
re
of mesopores. The BET surface area and total pore volume of UiO-66 activated by acetone
lP

via the Soxhlet extraction method for one day (AS-1) was measured to be 483 m2/g and

0.285 cm3/g, respectively, which are lower than the results reported for this MOF in the
na

literature[16, 24, 28]. With the increase of activation time, the BET surface areas decreased
ur

sharply to 23 and 29 m2/g for AS-5 and AS-10, respectively, reflecting the collapse of the
Jo

porous frameworks. A similar trend was also observed for the MOFs activated by acetone

via centrifugation method over different times. Therefore, these results indicated that the

activation of UiO-66 by acetone could only be carried out for one day in the Soxhlet

extraction method and five days in the centrifugation method. Above these cut-off times,

the porosity and crystallinity would be collapsed, thereby producing nonporous and

amorphous frameworks, which is consistent well with the XRD results. Moreover,

11
Journal Pre-proof

according to Tables 1 and S1, it can be found that the largest BET surface area and total

pore volume belonged to the UiO-66 activated by ethanol via the Soxhlet extraction method

for five days (ES-5), validating the well-preserved porosity and crystallinity of this MOF

under this condition.

of
ro
-p
re
lP
na
ur
Jo

12
Journal Pre-proof

of
ro
-p
re
lP
na
ur
Jo

13
Journal Pre-proof

Figure 2. N2 adsorption-desorption isotherms of UiO-66 activated via different solvents


(acetone, ethanol, and chloroform) and methods: (a) Soxhlet extraction and (b)
Centrifugation. A, C and E, as the first letter, denote acetone, chloroform, and ethanol,
respectively. S and C, as the second letter, signify Soxhlet extraction and centrifugation,
respectively.

The impact of the activation solvent and method on the thermal stability of the various

syntheized UiO-66s were investigated by thermogravimetric analyses (TGA). As shown in

of
Figure 3, the TGA curves of all synthesized UiO-66s showed a three-step weight loss,

ro
which are consistent with the literature [17, 29]. The first weight loss observed at

-p
temperature more than 100°C is related to the removal of adsorbed water molecules on the
re
pores and surface of UiO-66s. It was obvious that this weight loss for UiO-66 activated by

chloroform is significantly more than those of UiO-66s activated by acetone and ethanol.
lP

The second weight loss observed at temperature more than 200°C is associated to the
na

removal of DMF (solvent molecules) trapping in the pores of MOFs. The third weight loss

observed at temperature more than 500°C is attributed to the decomposition of UiO-66


ur

frameworks to CO2, CO, and ZrO2 [17]. The new weight loss ppeared around 400°C in the
Jo

TGA curve of AS-5 could be attributed to the decomposition of collapsed framework as

described in the XRD patterns of this UiO-66. Therefore, it can be expected that the

thermal stability of UiO-66 activated by acetone via Soxhlet extraction method is slightly

lower than those of activated by chloroform and ethanol.

14
Journal Pre-proof

of
ro
-p
re
lP
na

Figure 3. TGA (a and c) and derivative thermogravimetry curves (b and d) of UiO-66


ur

activated via different solvents (acetone, ethanol, and chloroform) and methods: (a) Soxhlet
extraction and (c) Centrifugation. A, C and E, as the first letter, denote acetone, chloroform,
and ethanol, respectively. S and C, as the second letter, signify Soxhlet extraction and
Jo

centrifugation, respectively.

The effect of the activation method and solvent on the morphology of the synthesized UiO-

66s was further studied by FESEM. Figure 4 depicts the FESEM images of UiO-66s

activated via different methods and solvents for five days. Moreover, for further

clarification, FESEM images of UiO-66s at various times are shown in Figure S4. It is clear

15
Journal Pre-proof

from the FESEM images that the crystals of UiO-66s are round-shaped with a

homogeneous morphology, which is in accord with the images reported in the literature

[17, 25, 27]. Additionally, it was found that changing the activation methods and solvents

resulted in insignificant changes in the morphology and particle size of the synthesized

UiO-66s.

of
ro
-p
re
lP
na
ur
Jo

Figure 4. FESEM images of the UiO-66s activated via the Soxhlet extraction and
centrifugation methods for five days. The scale bar is the same for all images.

3.2. Adsorption kinetics experiments

The time-dependent uptake capacity was obtained to investigate the kinetics for the

adsorption of MR and MB onto the synthesized MOFs, where the experimental is

16
Journal Pre-proof

delineated in the supporting information (Section S3). Adsorption kinetics of MR and MB

on the synthesized UiO-66s at room temperature is illustrated in Figure 5. It was unveiled

that increasing agitation time for all synthesized UiO-66s before achieving equilibrium led

to an increase in the uptake capacities of both dyes. As shown in Figure 5a, the uptake

capacity of MR onto all the synthesized UiO-66s presented nearly the same value, but the

equilibrium uptake capacity of MB onto these adsorbents is obviously different and

of
increased in the following order: CC-5 (77 mg/g) > CS-5 (60 mg/g) > EC-5 (42 mg/g) >

AC-5 (41 mg/g) > ES-5 (40 mg/g) > AS-5 (18 mg/g). This bodes a nearly three-time

ro
increase in the uptake capacity of MB for chloroform as opposed to acetone for activation.

-p
re
lP
na
ur
Jo

Figure 5. Effect of agitation time on the adsorption uptakes of (a) MR and (b) MB onto
UiO-66s.

ES-5 showed the highest uptake capacity for anionic MR dye (94.5 mg/g), which might be

17
Journal Pre-proof

due to its porous structure, large BET surface area, large total pore volume (Table S2). This

observation suggests that a large BET surface area and total pore volume increase the

uptake capacity of these adsorbents. However, Figure S6 speaks to the insensitivity of the

uptake capacities of these adsorbents to their BET surface areas. For instance, AS-5

presented the smallest BET surface area but showed a high uptake capacity for anionic MR

dye (91.6 mg/g). This unfolds that a high affinity of AS-5 towards anionic MR dye could

of
not be due to its porosity, and physical adsorption into its porous structure might not be the

only mechanism for its high uptake capacity. Therefore, we speculate that the high uptake

ro
capacity of AS-5 towards anionic MR dye could be related to its high positive zeta

-p
potential, as shown in Figure S7. Given this hypothesis, it can be asserted that the
re
electrostatic attraction between the positive charge of AS-5 and anionic MR molecules
lP

increased, resulting in an enhancement in the uptake capacity. Similar behavior was also

observed for the other synthesized adsorbents, where the uptake capacity of anionic MR
na

dye increased by the enhancement of the positive zeta potential.


ur

The adsorption kinetics of MR and MB on the UiO-66s activated by acetone for different
Jo

times was further studied (Figure S8). As presented, the uptake capacity of MR onto these

adsorbents relatively increased with increasing the activation time, while the uptake

capacity of MB showed an opposite trend. This observation can be attributed to the changes

in the zeta potential of adsorbents, as illustrated in Figure S7. Indeed, as a consequence of

changes in zeta potential, the electrostatic attraction between MR and adsorbents increased;

in contrast, the electrostatic attraction between MB and adsorbents decreased. This resulted

18
Journal Pre-proof

in an increase and a decrease in the uptake capacities of MR and MB, respectively.

Therefore, as shown in Figure S9, the ideal adsorption selectivity of MR over MB

improved by increasing the activation time. The details for calculating ideal adsorption

selectivity are provided in the supporting information (Section S4). Moreover, the ideal

adsorption selectivity of MR over MB onto the UiO-66s activated by different solvents and

various methods for five days was calculated, and the results are compiled in Table S3. As

of
presented in Table S3, the highest ideal adsorption selectivity (MR/MB) was observed for

AS-5, which could be attributed to its high positive zeta potential, as well as a high affinity

ro
towards anionic MR dye.

-p
To determine the rate-controlling step in the adsorption of MR and MB on the synthesized
re
UiO-66s, the pseudo-first-order, pseudo-second-order, and intra-particle diffusion kinetic
lP

models were used, where models are well described in the supporting information (Section

S3). All the calculated correlation coefficients (R2) of the pseudo-second-order model are
na

higher than those of the other models. Besides, it was observed that the calculated values of
ur

the uptake capacity (qe,cal) from the pseudo-second-order model for both dyes well match
Jo

the experimental data. These results endorse that the adsorption process of MR and MB

onto the synthesized adsorbents obeys the pseudo-second-order kinetic model [26, 30].

19
Journal Pre-proof

Table 1. Calculated kinetic parameters for the adsorption of MR and MB onto the
synthesized adsorbents.
R2
Adsorbents Dye Pseudo- Pseudo- Intra-particle qe (mg/g) K2 (g/mg min)
first-order second-order diffusion
MR 0.877 0.998 0.946 66.66 3.810-3
AS-1
MB 0.901 0.991 0.937 60.24 1.710-3
MR 0.959 0.999 0.944 71.42 4.610-3
AS-2
MB 0.590 0.995 0.863 14.74 1.310-2
MR 0.818 0.998 0.970 91.74 2.710-3
AS-3
4.610-3

of
MB 0.903 0.996 0.983 20.28
MR 0.867 0.995 0.919 46.08 3.710-3
AS-4
0.837 0.997 0.903 27.39 7.410-3

ro
MB
MR 0.751 0.999 0.931 92.59 5.510-3
AS-5
MB 0.929 0.986 0.964 19.26 3.010-3
AS-10
MR
MB
0.842
0.750
0.999
0.998 -p 0.974
0.974
91.74
13.38
3.610-3
2.010-2
re
MR 0.957 0.999 0.982 87.71 3.910-3
CS-5
MB 0.923 0.997 0.969 61.72 3.110-3
2.010-3
lP

MR 0.852 0.995 0.965 95.23


ES-5
MB 0.920 0.995 0.931 40.65 5.010-3
MR 0.937 0.998 0.976 97.08 1.610-3
AC-1
2.710-3
na

MB 0.938 0.995 0.949 51.81


MR 0.791 0.998 0.977 89.28 1.810-3
AC-2
MB 0.851 0.997 0.874 48.78 3.910-3
ur

MR 0.884 0.999 0.987 96.15 2.310-3


AC-3
MB 0.866 0.997 0.903 43.66 4.410-3
MR 0.842 0.989 0.944 85.47 9.410-4
Jo

AC-4
MB 0.845 0.998 0.944 47.16 5.010-3
MR 0.916 0.997 0.961 90.09 1.710-3
AC-5
MB 0.956 0.998 0.977 43.47 3.410-3
MR 0.818 0.996 0.956 95.23 1.410-3
AC-10
MB 0.855 0.996 0.902 50.25 3.610-3
MR 0.948 0.999 0.977 94.33 2.510-3
CC-5
MB 0.946 0.997 0.978 80.00 1.810-3
MR 0.961 0.998 0.958 85.47 2.810-3
EC-5
MB 0.924 0.996 0.963 44.05 3.510-3

20
Journal Pre-proof

3.3. Equilibrium adsorption isotherms

It is clear that the uptake capacities of different dyes are influenced by their initial

concentrations. Therefore, the maximum uptake capacities of MR and MB on ES-5, as the

representative of all synthesized adsorbents, were examined by testing their uptake

capacities at various initial dye concentrations. Among the different synthesized samples,

ES-5 was chosen because of its high BET surface area, high crystallinity, and high uptake

of
capacity for MR dye. As presented in Figure 6, the maximum uptake capacity for anionic

ro
MR dye and cationic MB dye is equal to 243 and 217 mg/g, respectively, which are slightly

-p
higher than those reported in the literature (Table S4) [16, 31-33].
re
lP
na
ur
Jo

Figure 6. Equilibrium adsorption isotherms of MR and MB onto UiO-66 activated by


ethanol via the Soxhlet extraction (ES-5) method.

21
Journal Pre-proof

To further determine the adsorption mechanisms, surface properties of the adsorbents, and

the affinity of the adsorbents to the adsorbates, four adsorption isotherm models, including

Langmuir, Freundlich, Dubinin-Radushkevich, and Tempkin isotherms, were used and

fitted to the experimental adsorption data. The details of these isotherm models are

presented in the supporting information (Section S5). As shown in Figure S10 and Table 2,

the Langmuir isotherm model has higher correlation coefficients (R2) than those of the

of
other mentioned isotherm models. These results indicate that the adsorption of MR and MB

dyes onto ES-5 obeyed the Langmuir isotherm model, signifying a monolayer adsorption

ro
process [34]. Moreover, the calculated standard free energy change (∆G0) values for MR

-p
and MB dyes are negative, revealing that the adsorption of these dyes onto ES-5 is a
re
spontaneous adsorption thermodynamic process in the studied experiments (Table S5). The
lP

thermodynamic investigation is explained in detail in the supporting information (Section

S6).
na
ur

Table 2. Constant isotherm parameters of MR and MB onto UiO-66 activated by ethanol


Jo

via the Soxhlet extraction (ES-5) method.


Langmuir constants Freundlich constants Dubinin-Radushkevich constants Tempkin constants
Dyes
qmax KL KF qmax
R2 n R2 B E (J/mol) R2 BT AT R2
(mg/g) (L/mg) (mg/g) (mg/g)

MR 243.90 0.1297 0.987 87.19 5.01 0.996 176.72 310-8 4.1103 0.742 103.55 69.02 0.952

MB 217.39 0.0428 0.999 19.63 2.07 0.951 142.43 410-6 3.5102 0.775 56.37 0.512 0.987

22
Journal Pre-proof

3.4. Effect of pH

The pH of the dye solutions is one of the governing factors for the uptake capacity in the

adsorption process as it may affect the zeta potential of adsorbents, the ionization degree,

and the chemical stability of organic dyes. Therefore, to investigate the effect of solution

pH on the uptake capacities of MR and MB dyes, the adsorption experiments were

performed in the solution pH range of 2.5-13.1. The initial pH of the solutions was adjusted

of
by using diluted HCl and NaOH solutions. As shown in Figure 7a, the uptake capacity of

ro
anionic MR dye was high at low pH values, which is due to the strong electrostatic

-p
attraction with the ES-5 surface. However, it can be seen that the uptake capacity of MR

dye declined slowly until pH 5.8, and then decreased drastically with increasing the pH up
re
to 13. This reduction in the uptake capacity at higher pH values might be due to an increase
lP

in electrostatic repulsion between the adsorbent and the adsorbate as a result of changes in
na

the zeta potential of adsorbent nanoparticles (Figure 7b) [35].

The adsorption performance of ES-5 for cationic MB dye at various pH values showed a
ur

strong pH-dependent behavior. Accordingly, at low pH values, there is strong electrostatic


Jo

repulsion between cationic MB molecules and highly positive charged ES-5 nanoparticles,

as well as intense competition between MB molecules and H3O+ ions to be adsorbed onto

the ES-5 nanoparticles. Therefore, as presented in Figure 6a, the uptake capacity of MB dye

at low pH values was small. With an increase in solution pH, the zeta potential of adsorbent

decreased, and the electrostatic repulsion between the cationic MB dye and the positive

adsorbent started to decline, and electrostatic attraction increased. This resulted in an

23
Journal Pre-proof

increase in the uptake capacity of MB dye with an increase in solution pH, reaching a

maximum value at pH 9.1. However, at higher pH values, the uptake capacity of MB

decreased, which might be due to the conversion of MB molecules into other compounds

via hydrolysis in strongly basic solutions, and partial collapse of the ES-5 framework in a

basic solution as shown in the XRD patterns of ES-5 after being stirred in basic solution for

2 h (Figure 7c) [36]. While, the XRD patterns of this MOF after being stirred in acidic

of
solution don’t show any change. Therefor, it can be said that the crystalline structure of ES-

ro
5 is stable in acidic solution, while it partially collapsed at basic solution.

-p
Since the pH of the solution is an essential parameter for changing the surface charge of

adsorbents, the zeta potential of ES-5 was determined in the pH range of 2-13 (Figure 7b).
re
The zeta potential of ES-5 nanoparticles decreased drastically with an increase of the
lP

solution pH, suggesting that the zeta potential of this adsorbent is highly pH-dependent.
na

Additionally, as a consequence of changes in the zeta potential of adsorbent nanoparticles

as well as changes in the interactions between the adsorbent and the adsorbates, the ideal
ur

adsorption selectivity of anionic MR dye over cationic MB dye decreased with raising the
Jo

solution pH (Figure 7a).

24
Journal Pre-proof

of
Figure 7. (a) Effect of pH on the adsorption uptake of MR and MB onto ES-5, (b) the zeta
potential of ES-5 as a function of pH, and (c) XRD patterns of ES-5 before and after stirred

ro
in acidic (pH≈2.2) and basic (pH≈12.8) solutions for 2 h.

3.5. Adsorption mechanism


-p
re
According to the results obtained from zeta potential measurement and pH effect, it can be
lP

said that the electrostatic interaction of anionic MR and cationic MB dyes with charged
na

nanoparticles of ES-5 is among the main adsorption mechanisms. However, to illuminate

the detailed adsorption mechanism of ES-5 towards MR and MB dyes, the FTIR assay was
ur

carried out to detect the changes in the FTIR spectrum of ES-5 before and after adsorption
Jo

of these dyes (Figure 8). FTIR spectrum of SE-5 shows some notable redshift after MR dye

adsorption, which are assigned to the carboxyl and hydroxyl stretching vibration bands.

This suggests the formation of hydrogen-bonding interactions between ES-5 and MR

molecules [2, 37]. Moreover, it is evident that the aromatic ring vibration of both MR and

MB dyes decreased after being adsorbed onto the surface of ES-5 nanoparticles. This

illustrates that the π-π stacking interactions between the aromatic portion of ES-5 and the

25
Journal Pre-proof

aromatic rings of dye molecules might be another mechanism for the excellent adsorption

performance of ES-5 towards these dyes [1, 38].

of
ro
-p
re
lP
na

Figure 8. FTIR spectra of pure dyes and ES-5 before and after adsorption of (a) MR and
(b) MB.
ur
Jo

3.6. Regeneration and reusability of ES-5

For commercial adsorbents, regeneration and reusability are of great importance. Therefore,

regeneration and reusability of dyes-loaded ES-5 are evaluated using acidic ethanol as an

eluent. Accordingly, the dyes-loaded nanoparticles were dispersed in acidic ethanol

solution by an ultrasonic assistant for 60 min, washed three times with fresh ethanol and

further with water, dried at 120°C under vacuum for 12 h, and then applied for further dye

26
Journal Pre-proof

adsorption process. As depicted in Figure 9, the adsorption capacity of ES-5 for both

cationic and anionic dyes after five consecutive adsorption-desorption cycle show a

negligible reduction, which demonstrates good regeneration and reusability of ES-5.

of
ro
-p
re
lP
na

Figure 9. Regeneration and reusability of ES-5.


ur
Jo

4. Conclusions

In the present work, a simple method was developed to activate UiO-66 with adjustable

specific surface areas and pore volumes via the centrifugation and Soxhlet extraction

methods. The different synthesized UiO-66s were then employed to remove anionic methyl

red (MR) and cationic methylene blue (MB) dyes from aqueous solutions. Among the

27
Journal Pre-proof

different synthesized UiO-66s, ES-5 showed the highest uptake capacity for MR (243

mg/g) and MB (217 mg/g) dyes. These high uptake capacities might derive from strong

electrostatic interactions, π-π stacking interactions, and hydrogen bonding between

adsorbent nanoparticles and dye molecules. In addition, the adsorption isotherms and

adsorption kinetics could be well described by the Langmuir isotherm model and pseudo-

second-order kinetic model, respectively. Besides, the thermodynamic investigation

of
showed that the adsorption of both dyes is spontaneous and endothermic. The findings of

this study indicate that the uptake capacity of organic dyes not only depends on surface area

ro
and pore volume of adsorbents but also depends on the surface charge of adsorbents.

-p
re
lP

Acknowledgments

The authors acknowledge the financial support received from the University of British
na

Columbia, Okanagan Campus. Dr. Arjmand would like to thank the financial support from
ur

the Canada Research Chairs program.


Jo

5. References

[1] H. Molavi, M. Neshastehgar, A. Shojaei, H. Ghashghaeinejad, Chemosphere 247 (2020)


125882.
[2] M. Neshastehgar, P. Rahmani, A. Shojaei, H. Molavi, Microporous and Mesoporous
Materials 296 (2020) 110008.

28
Journal Pre-proof

[3] J. Chen, Y. Sheng, Y. Song, M. Chang, X. Zhang, L. Cui, D. Meng, H. Zhu, Z. Shi, H.
Zou, ACS Sustainable Chemistry & Engineering 6 (2018) 3533.
[4] M. Sheikh, M. Pazirofteh, M. Dehghani, M. Asghari, M. Rezakazemi, C. Valderrama,
J.-L. Cortina, Chemical Engineering Journal 391 (2020) 123475.
[5] P. Hu, Z. Zhao, X. Sun, Y. Muhammad, J. Li, S. Chen, C. Pang, T. Liao, Z. Zhao,
Chemical Engineering Journal 356 (2019) 329.
[6] Y. Duan, F. Ye, Y. Huang, Y. Qin, C. He, S. Zhao, Chemical Communications 54
(2018) 5377.
[7] M. Zheng, S. Liu, X. Guan, Z. Xie, ACS Applied Materials & Interfaces 7 (2015)

of
22181.
[8] H. Molavi, F.A. Joukani, A. Shojaei, Industrial & Engineering Chemistry Research 57

ro
(2018) 7030.
[9] H. Molavi, H. Moghimi, R.A. Taheri, Applied Organometallic Chemistry 34 (2020)
e5549.
-p
re
[10] M. Rezakazemi, M. Sadrzadeh, T. Matsuura, Progress in Energy and Combustion
Science 66 (2018) 1.
lP

[11] M. Rezakazemi, A. Ebadi Amooghin, M.M. Montazer-Rahmati, A.F. Ismail, T.


Matsuura, Progress in Polymer Science 39 (2014) 817.
[12] Z. Li, W. Sun, C. Chen, Q. Guo, X. Li, M. Gu, N. Feng, J. Ding, H. Wan, G. Guan,
na

Applied Surface Science 480 (2019) 770.


[13] S.A. Sadat, A.M. Ghaedi, M. Panahimehr, M.M. Baneshi, A. Vafaei, M. Ansarizadeh,
ur

Applied Surface Science 467 (2019) 1204.


[14] H. Mahdavi, L. Ahmadian‐ Alam, H. Molavi, Polymer International 64 (2015) 1578.
Jo

[15] M. Gomar, S. Yeganegi, Microporous and Mesoporous Materials 252 (2017) 167.
[16] Q. Chen, Q. He, M. Lv, Y. Xu, H. Yang, X. Liu, F. Wei, Applied Surface Science 327
(2015) 77.
[17] H. Molavi, A. Hakimian, A. Shojaei, M. Raeiszadeh, Applied Surface Science 445
(2018) 424.
[18] M.I. Hossain, T.G. Glover, Industrial & Engineering Chemistry Research 58 (2019)
10550.
[19] F. Ahmadijokani, R. Mohammadkhani, S. Ahmadipouya, A. Shokrgozar, M.

29
Journal Pre-proof

Rezakazemi, H. Molavi, T.M. Aminabhavi, M. Arjmand, Chemical Engineering Journal


399 (2020) 125346.
[20] Z. Tahir, M. Aslam, M.A. Gilani, M.R. Bilad, M.W. Anjum, L.-P. Zhu, A.L. Khan,
Separation and Purification Technology 224 (2019) 524.
[21] M. Younas, M. Rezakazemi, M. Daud, M.B. Wazir, S. Ahmad, N. Ullah, Inamuddin,
S. Ramakrishna, Progress in Energy and Combustion Science 80 (2020) 100849.
[22] H.R. Abid, H. Tian, H.-M. Ang, M.O. Tade, C.E. Buckley, S. Wang, Chemical
Engineering Journal 187 (2012) 415.
[23] S. Tai, W. Zhang, J. Zhang, G. Luo, Y. Jia, M. Deng, Y. Ling, Microporous and

of
Mesoporous Materials 220 (2016) 148.
[24] H.R. Abid, J. Shang, H.-M. Ang, S. Wang, International Journal of Smart and Nano

ro
Materials 4 (2013) 72.
[25] S. Edubilli, S. Gumma, Separation and Purification Technology 224 (2019) 85.
-p
[26] J. Qiu, Y. Feng, X. Zhang, M. Jia, J. Yao, Journal of colloid and interface science 499
re
(2017) 151.
[27] J.H. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, S. Bordiga, K.P.
lP

Lillerud, Journal of the American Chemical Society 130 (2008) 13850.


[28] H.R. Abid, G.H. Pham, H.-M. Ang, M.O. Tade, S. Wang, Journal of colloid and
interface science 366 (2012) 120.
na

[29] H. Molavi, M. Zamani, M. Aghajanzadeh, H. Kheiri Manjili, H. Danafar, A. Shojaei,


Applied Organometallic Chemistry 32 (2018) e4221.
ur

[30] W. Konicki, M. Aleksandrzak, D. Moszyński, E. Mijowska, Journal of colloid and


interface science 496 (2017) 188.
Jo

[31] E. Haque, V. Lo, A.I. Minett, A.T. Harris, T.L. Church, Journal of Materials
Chemistry A 2 (2014) 193.
[32] Z. Ioannou, C. Karasavvidis, A. Dimirkou, V. Antoniadis, Water Science and
Technology 67 (2013) 1129.
[33] T. Santhi, S. Manonmani, T. Smitha, Chemical engineering research bulletin 14 (2010)
11.
[34] Y. Li, Q. Du, T. Liu, X. Peng, J. Wang, J. Sun, Y. Wang, S. Wu, Z. Wang, Y. Xia,
Chemical Engineering Research and Design 91 (2013) 361.

30
Journal Pre-proof

[35] J. Yan, G. Lan, H. Qiu, C. Chen, Y. Liu, G. Du, J. Zhang, Separation Science and
Technology 53 (2018) 1678.
[36] C. Li, Z. Xiong, J. Zhang, C. Wu, Journal of Chemical & Engineering Data 60 (2015)
3414.
[37] M. Tong, D. Liu, Q. Yang, S. Devautour-Vinot, G. Maurin, C. Zhong, Journal of
Materials Chemistry A 1 (2013) 8534.
[38] J.-J. Li, C.-C. Wang, H.-f. Fu, J.-R. Cui, P. Xu, J. Guo, J.-R. Li, Dalton transactions 46
(2017) 10197.

of
ro
-p
re
lP
na
ur
Jo

31
Journal Pre-proof

Conflict of Interest

The authors declare that there is no conflict of interest.

of
ro
-p
re
lP
na
ur
Jo

32
Journal Pre-proof

Credit Author Statement

S. No. First Name, Sur Name Role

Danial Vaghar Investigation, Validation, Formal analysis, Data


1
Mousavi Curation

of
Investigation, Validation, Formal analysis, Data

ro
2. Salman Ahmadipouya
Curation

3. Atefeh Shokrgozar
-p
Investigation, Validation, Formal analysis
re
Conceptualization, Methodology, Project administration,
lP

4. Hossein Molavi
Supervision, Writing - Original Draft
na

Mashallah
5. Writing - Review & Editing
ur

Rezakazemi
Jo

6. Farhad Ahmadijokani Investigation, Formal analysis, Review & Editing

Conceptualization, Writing - Review & Editing,


7. Mohammad Arjmand
Supervision, Funding acquisition

33
Journal Pre-proof

Research Highlights

UiO-66 was activated via the centrifugation and Soxhlet extraction methods.

UiO-66 was activated by acetone, chloroform, and ethanol solvents.

Among the three activation solvents, ethanol presents the best performance for activation of

UiO-66.

of
UiO-66 activated by ethanol via the Soxhlet extraction method shows the highest uptake

ro
capacity for methyl red (243 mg/g) and methylene blue (217 mg/g) dyes.

-p
re
lP
na
ur
Jo

34

You might also like