You are on page 1of 36

ChemistrySelect

A new resin containing aminopropylphosphonate chelating ligand for high-performance


mitigation of heavy metal ions
--Manuscript Draft--

Manuscript Number: slct.201803045

Article Type: Full Paper

Corresponding Author: Izzat W Kazi, Ph.D.


King Fahd University of Petroleum & Minerals
Dhahran, SAUDI ARABIA

Corresponding Author E-Mail: iwkazi@kfupm.edu.sa

Order of Authors (with Contributor Roles): Izzat W Kazi, Ph.D.

Nisar Ullah, Ph.D.

Shaikh Asrof Ali, Ph.D.

Keywords: Cyclopolymerization; adsorption; chelating resin; aminopropylphosphonate; heavy


metal ions

Manuscript Classifications: Environmental chemistry; Water chemistry

Suggested Reviewers: Osman S Kabasakal, Ph.D.


Eskisehir Osmangazi Univ Turkey
osk@ogu.edu.tr
Expert in metal ion removal

Fenglian Fu, Ph.D.


Guangdong Univ of Technology China
Fufenglian2006@163.com
Expert in metal ion removal

Ali Duran
Hacettepe Univ Turkey
ali.duran@kosgeb.gov.tr
Expert in metal ion removal

Junsheng Liu
Hefei Univ China
jsliu@hfuu.edu.cn
Expert in metal ion removal

Opposed Reviewers:

Abstract: Cyclopolymerization, initiated by t-butylhydroperoxide, of an aqueous solution of N-(3-


phosphonopropyl)diallylammonium hydrochloride and cross-linker
tetraallylpiperazinium dichloride in 9:1 mol ratio, led to a pH-responsive cross-linked
resin (80% yield). The resin showed exceptional efficiency for entrapping toxic metal
ions from aqueous systems. Using Langmuir adsorption isotherm, the respective
maximum uptake (Qm) of Pb2+ and Cu2+ ions was determined to be 7.18 and 6.76
mmol g-1. The very impressive performance accorded the resin a prestigious place
among many sorbents in recent works. The adsorption data fitted second-order
kinetics with respective Ea of 35.5 and 31.1 kJ mol-1 for Pb2+and Cu2+ ions. While the
initial rapid adsorption of metal ions was due to film diffusion, the slower adsorption
followed intraparticle diffusion. SEM and EDX analyses confirmed the metal ions’
adsorption on to sorbent. The negative Gº and positive Hº values demonstrated the
endothermic adsorption as a favorable process. The sorbent is recyclable as evinced
by the excellent adsorption/desorption process.

Author Comments: No comment

Section/Category: Materials Science inc. Nanomaterials & Polymers

Additional Information:

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Question Response

Submitted solely to this journal? Yes

Has there been a previous version? No

Do you or any of your co-authors have a No. The authors declare no conflict of interest.
conflict of interest to declare?

Animal/tissue experiments?

Powered by Editorial Manager® and ProduXion Manager® from Aries Systems Corporation
Manuscript Click here to download Manuscript MAUSCRIPT.docx

1
2
3
4
5 A new resin containing aminopropylphosphonate
6
7
8 chelating ligand for high-performance mitigation of
9
10
11
12
heavy metal ions
13
14
15
16
17
Izzat W. Kazi*● Nisar Ullah● Shaikh A. Ali
18
19
20
21
22 Chemistry Department, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi
23
24 Arabia
25
26
27
28
29
30
31 *Corresponding author
32
33
34 E-mail address:iwkazi@kfupm.edu.sa
35
36
37 Tel.: (966)13 860 8328; Fax: (966) 13 860 4277
38
39
40 Home Page: http://faculty.kfupm.edu.sa/CHEM/iwkazi/
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 1
65
1
2
3
4 Abstract Cyclopolymerization, initiated by t-butylhydroperoxide, of an aqueous solution of N-(3-
5
6
7 phosphonopropyl)diallylammonium hydrochloride and cross-linker tetraallylpiperazinium
8
9 dichloride in 9:1 mol ratio, led to a pH-responsive cross-linked resin (80% yield). The resin showed
10
11
12 exceptional efficiency for entrapping toxic metal ions from aqueous systems. Using Langmuir
13
14 adsorption isotherm, the respective maximum uptake (Qm) of Pb2+ and Cu2+ ions was determined
15
16
17 to be 7.18 and 6.76 mmol g-1. The very impressive performance accorded the resin a prestigious
18
19 place among many sorbents in recent works. The adsorption data fitted second-order kinetics with
20
21 respective Ea of 35.5 and 31.1 kJ mol-1 for Pb2+and Cu2+ ions. While the initial rapid adsorption of
22
23
24 metal ions was due to film diffusion, the slower adsorption followed intraparticle diffusion. SEM
25
26 and EDX analyses confirmed the metal ions’ adsorption on to sorbent. The negative Gº and
27
28
29 positive Hº values demonstrated the endothermic adsorption as a favorable process. The sorbent
30
31
32
is recyclable as evinced by the excellent adsorption/desorption process.
33
34
35
36 Keywords: Cyclopolymerization ● adsorption ● chelating resin ● desorption ●

37
38 aminopropylphosphonate heavy metal ions wastewater
● ●

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
64 2
65
1
2
3
4
5 1 INTRODUCTION
6
7
8 Sustenance of human civilization demands clean environment where water quality is the single
9
10 most important issue. The indiscriminate disposal of industrial effluents from manufacturing
11
12
13 industries such as refinery, paint, metallurgical, mining, chemical, battery, etc. has led to
14
15 recalcitrant nonbiodegradable toxic heavy metal pollution, which has become a global epidemic
16
17
18 and thus captured attention from the research community (Ngah and Hanafiah 2008). Chronic
19
20 exposure to lead, contained in wastewaters from various industries, is so detrimental to health of
21
22
23
human and animals that the U.S. Environmental Protection Agency (EPA) has set a lead
24
25 concentration limit approaching zero for drinking water (US EPA 2009), whereas the Chinese
26
27 standard (Ministry of Health 2006) has set 0.01 mg L-1 (i.e. 0.01 ppm or 10 ppb) as the extremely
28
29
30 severe contaminant level. Adsorption as a technique for heavy metal removal offers numerous
31
32 advantages owing to its lower cost and higher efficiency, lower reagent or energy requirement (Fu
33
34
35 and Wang 2011; Mishra 2014; Hua et al. 2012). For the mitigation of Pb(II) contamination, various
36
37 adsorbents have been developed (Ahmad et al. 2017; Marzougui et al. 2016; Xie et al. 2015; Ray
38
39
40 and Shipley 2015).
41
42 Copper in a small amount is an essential nutrient. However, above prescribed limit, it can cause
43
44
45
gastrointestinal disturbance, liver or kidney damage. The EPA has set a maximum contaminant
46
47 level at 1.3 parts per million (ppm) for copper in drinking water. By virtue of a natural mechanism,
48
49 the human body can maintain the proper level of copper in it. However, children, yet to develop
50
51
52 this mechanism, are more vulnerable to the toxic effects of copper. Various adsorbents (He et al.
53
54 2016; Shen et al. 2017), including magnetic chitosan nanoparticles (Yuwei and Jianlong 2011) and
55
56
57 amino-functionalized magnetic nanoparticles (Hao et al. 2010), have been used to entrap Cu2+
58
59 from aqueous systems.
60
61
62
63
64 3
65
1
2
3
4 Currently, there is a trend of using graphene decorated with functionalities to trap toxic metal
5
6
7 ions (Yusuf et al. 2015; Rao et al. 2017). A recent review describes the use of functionalized
8
9 polymer-nanocomposites for metals removal from contaminated aqueous systems (Lofrano et al.
10
11
12 2016).
13
14 In ion exchange resins having covalently attached X- (SO3-, PO32- or CO2-) functionalities, the
15
16
17 counterions like H+ or Na+ are exchanged with toxic metal ions Mn+. Amphoteric exchangers, on
18
19 the other hand, contains zwitterionic dipole: whose charges of both algebraic signs offers a greater
20
21 scope of entrapping both cations or anions (Liu et al. 2010; Wang et al. 2014; Samiey et al. 2014).
22
23
24 Chelating ion exchangers that contain chelating ligands are becoming increasingly attractive for
25
26 metal ion removal (Alexandratos 2009; Ali et al. 2015; Jiang et al. 2015).
27
28
29 Aminomethylphosphonate and aspartate functional motifs has been remarkably selective towards
30
31 heavy metal ions (Jamiu et al. 2015; Deepatana and Valix 2006).
32
33
34 Industrial wastewaters need to be recycled after remediation of harmful metals (Ranade and
35
36 Bhandari 2014). The requirement of fast entrapment of metal ions demands exploration of new
37
38
39
adsorbents. The current article describes the synthesis of a new pH-responsive ionic resin
40
41 embedded with aminopropylphosphonate as chelating motifs for scavenging Pb2+ and Cu2+ ions
42
43 removal from aqueous systems. (Scheme 1).
44
45
46
47 2 MATERIALS AND METHODS
48
49
50
51 2.1 Physical Methods
52
53
54
55 IR spectra were recorded in a Nicolet 6700 FTIR spectrometer (Thermo Electron Corporation,
56
57 USA), while atomic compositions were determined by a Perkin Elmer Elemental Analyzer 2400
58
59
60 (Waltham, Massachusetts, USA). SEM and EDX were obtained in a TESCAN LYRA 3 (Czech
61
62
63
64 4
65
1
2
3
4 Republic) instrument. With a temperature increase by 15 °C/min and N2 flow of 50 cm3/min, TGA
5
6
7 was recorded in an SDT analyzer (Q600: TA instruments, USA) in the range 20–800 °C. Brunauer-
8
9 Emmett-Teller (BET) method was utilized to determine surface area and pore size. XRD analysis
10
11
12 was performed using a MiniFlex II diffractometer (Rigaku model, Japan). The metal ions
13
14 concentrations in wastewater were found using ICP-MS (ICP-MS XSERIES-II Thermo
15
16
17 Scientific).
18
19
20 2.2 Materials
21
22
23
24 Tert-butylhydroperoxide (TBHP) was purchased from Fluka AG. Monomer 1 (Ali et al. 2015) and
25
26
27
cross-linker 2 (Ali et al. 1996) were prepared as described.
28
29
30 2.3 Synthesis of crosslinked resins
31
32
33
34 2.3.1 Synthesis of Cross-linked polyzwitterion (CPZ) 3
35
36
37 A solution of 1 (10.2 g, 40 mmol), 2 (1.42 g, 4.44 mmol), water (5.0 g) (i.e. the monomers/water
38
39
40 at a 70:30 wt ratio) and the initiator TBHP (970 mg) in a 50 mL RB flask under N2 was stirred at
41
42 85 °C for 4 h. The solution was transformed into a gel. The continued stirring at 95 °C for a further
43
44
45 24 h led to resin CPZ 3 which was washed with water and vacuum dried at 70 °C (8.2 g, 80%).
46
47 CPZ 3 contained 1 (-HCl): C9H18NO3P (90 mol%) and 2: C16H28Cl2N2 (10 mol%) (Found C, 50.6;
48
49
50 H, 8.6; N, 6.4 % requires C, 50.83; H, 8.35; N, 6.72%).
51
52
53
54
55
56
57
58
59
60
61
62
63
64 5
65
1
2
3
4 2.3.2 Transformation of CPZ 3 to cross-linked polydianion (CPD) 4
5
6
7
8 Resin 3 (7.0 g 30.5 mmol) was stirred with NaOH (2.8 g, 70 mmol) in aqueous solution (100 mL)
9
10 for 1 h at 20 °C. Methanol (100 mL)/NaOH (1 g, 25 mmol) solution was added to the resultant
11
12
13 CPD 4 followed by acetone (200 mL). CPD 4 was filtered, washed with acetone, and vacuum dried
14
15 at 70 °C (7.2 g, 89%). (Found: C, 43.2; H, 6.8; N, 5.6%. Repeating unit, taken as
16
17
(C9H16NNa2O3P)0.9 (C16H30N2O2)0.1, Requires C, 43.95; H, 6.62; N, 5.81%).
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
64 6
65
1
2
3
4
5 i) (CH3)3OOH
6
7 +
8 N Cl N Cl ii) soaked in H2O
9 H (CH2)3 (-HCl)
10 HO
11 P O
12 HO N Cl
13
14 1
15
16
17 2
18
19
20 [ ] [ ] [ ]
21 pn 0.10n (0.90-p)n
22
23
24 N N Cl N
25
H (CH2)3 H (CH2)3
26
O O
P O P O
27 HO N Cl HO
28
29
30
31
32
33
3 Crosslinked polyzwitterion (CPZ)
34
35
36 NaOH
37
38 [ ] [ ] [ ]
39 pn 0.10n (0.90-p)n
40
41
42 N N HO N
43 (CH2)3 (CH2)3
44 Na+ -O Na+ -O
P O P O
45 Na+ -O N HO Na+ -O
46
47
48
49
50
51 4 Crosslinked polydianion (CPD)
52
53
54
55
56 Scheme
Scheme 1. 1. Synthesis
Synthesis of of cross-linked
resin cyclopolymer containing
having aminopropylphosphonate residues
57 aminopropylphosphonate residues.
58
59
60
61
62
63
64 7
65
1
2
3
4
5
2.4 Ion exchange capacity (IEC)
6
7
8
9 Resin 4 (100 mg) was immersed in 0.1 M HCl (50 mL) for overnight, thereafter IEC in mmol g-
10
1
11 was determined from the excess acid in the supernatant by titrating with 0.1 M NaOH using eq
12
13
14
(1):
15
16 ( M i  M f ) 50
17 IEC  (1)
18
W
19
20
21
22 where Mi and Mf represent the molarity of HCl in the initial and final solutions, and W in g denotes
23
24 the polymer mass.
25
26
27
28
29 2.5 Adsorption capacities
30
31
32
33 In acetate buffer of variable pH, resin 4 (50 mg) and 0.1 M Pb(NO3)2 (20 mL) was stirred (24 h).
34
35 As described by Kayal and Singh 2007, the concentration of Pb2+ left-over in the combined filtrate
36
37
38 was quantified by titration with 0.01-0.1 M EDTA solution. Describing Co and Cf as Pb2+ ions’ the
39
40 respective initial and final molarity (M); V as the volume in mL and and W as resin mass in g, the
41
42
43 adsorption capacity ( q Pb2 , mmol g-1) was determined from Eq. (2):
44
45
46
47 (C0  Cf ) V
48 qPb2  (2)
49 W
50
51 The above titration using solutions of known concentrations (0.01-0.1 M) of Pb2+ were accurate to
52
53 ±1.5%. Triplicate adsorption data varied in the range 1-2.3%.
54
55
56
57
58
59
60
61
62
63
64 8
65
1
2
3
4 The concentrations of Cu2+ ions in the stock solutions of Cu(NO3)2 (0.1-0.01 M) and test
5
6
7 solutions, carried out using iodometric titration method (Liu et al. 2010), were found to be accurate
8
9 to ±2% and varied in the range 1-3%, respectively.
10
11
12 For kinetic-runs at different temperatures, resin 4 (250 mg)/0.1 M M2+ (100 mL) mixture at pH
13
14 5.3 was stirred, and the metal ions concentrations versus time were determined by titrating a small
15
16
17 volume of filtered aliquots. Using concentrations in the range 0.02 - 0.1 M, the adsorption
18
19 capacities of the resin were determined to fit several isotherms and determine thermodynamic
20
21
22
parameters Go, Ho and So.
23
24
25 2.6 Desorption process
26
27
28
29 After stirring resin 4 (50 mg)/0.1 M M2+ (20 mL) mixture at pH 5.3 for 24 h, it was centrifuged;
30
31
32 the qe values were calculated using the supernatant. The metal-loaded resin in the centrifuge tube
33
34 was washed with a pH 5.3 buffer and then treated with 0.5 M HNO3 (50 mL) at 20 °C for 2 h. The
35
36
desorption efficiency was calculated using the concentration of desorbed metal ions. The
37
38
39 adsorption/desorption procedure was repeated three times with the desorbed resin.
40
41
42
43 3 RESULTS AND DISCUSSION
44
45
46 3.1 Resin Synthesis
47
48
49
50 Utilizing Butler’s protocol (Butler 2000; Kudaibergenov et al. 2006; Singh et al. 2007), monomers
51
52
53 1 and 2 underwent polymerization to give CPZ 3 in 80% yield (Scheme 1). The protocol (Butler
54
55 2000) leads to the eighth most significant architecture (McGrew 1958) containing pyrrolidine rings
56
57
in the macromolecule. Monomer 2 having four allyl motifs led to the formation of cross-linked
58
59
60 resin, and at high conversion (80%), 1/2 composition is expected to be 9:1 (same as feed ratio) as
61
62
63
64 9
65
1
2
3
4 ascertained by elemental analysis. CPZ 3 was treated with NaOH to generate CPD 4 in excellent
5
6
7 yield (89%). Note that monomer 1 can act as chelating ligands as it has three ligand centers: one
8
9 in N and two in PO32- which can act.
10
11
12 The surface area of resin 4 was found to be 1.72 m2 g−1. The TGA curve of CDAP 4 in Fig.
13
14 1a shows two major accelerated weight losses of 20 and 41% in 25-200 and 400-500 °C ranges,
15
16
17 respectively. The first one could account for the loss of hydrated water, while the second major
18
19 loss indicates the loss of diallylamine moiety. The residual mass of 25% may belong to the sodium
20
21 phosphonate fraction.
22
23
24
25 3.2 Adsorption versus pH
26
27
28
29 Equilibrium q M 2  was determined at pH range 2.8-5.3 using acetate buffer. The qCu 2 increased
30
31
32
33
with the increase in pH values, while for Pb2+ ions, q Pb2 value is found to attain a maximum at a
34
35 pH of 3.4 and then passes through a minimum at pH 4.6 (Fig. 1b). Note that the adsorption
36
37
38 capacities for both the ions are very low at a pH of 2.8, thereby suggesting that the desorption and
39
40 adsorption process should be effective at lower and higher pH values, respectively. A pH of 5.3
41
42
43 was maintained for kinetic and thermodynamic studies; at the higher pH values the metal
44
45 hydroxide precipitation, which could lead to erroneous values for the q M 2  values (Veli and Pekey
46
47
48 2004). At higher values of pH, the short supply of H+ ions, which competes for the adsorption sites,
49
50
51 favors the uptake of metal ions.
52
53
54
55
56
57
58
59
60
61
62
63
64 10
65
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
Fig. 1 (a) TGA curve of CDP 4, (b) Plots of q M 2  - pH at 295 K for initial metal
25 concentration of 0.1 M.
26
27
28
29 The resin, at higher pH values, has higher negative charge density which promotes metal ion
30
31 exchange as well as chelation to its ligand centers (Scheme 2) (Kołoynska et al. 2008). At higher
32
33
34 metal ion concentrations, the qm2+ values could be enhanced by switching the bidentate ligand as
35
36 depicted in B to capture two M2+ ions by ion exchange (Scheme 2).
37
38
39
40 3.3 IEC of CPD 4 and its IR spectrum
41
42
43
44 The IEC of CPD 4 was determined to have a high value of 5.18 mmol g-1 which reflects its
45
46 capability to have higher uptake of metal ions. Both ion exchange and chelation can lead to the
47
48
49 chemical adsorption (Scheme 2).
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 11
65
1
2
3
4
5
6 N 2+
N M ON
7 H OM2+ NO - H
8 HO 3 O
P O O P
9 P
10 O M 2+
11 O O
12
13 pKa A B C
N M 2+
14 X Monodentate Bidentate Tridentate
15 10.6 H
16 7 HO P
17
2.4 HO
18 O
19
20 N N NM2+
21
H O O H O
HO OH HO
22 P M 2+ P P
23
24 O O O
25
26 D E
27
28 Scheme 2. Possible modes of capturing metal ions by aminopropylphosphonate
29 chelating ligands.
30
31
Scheme 2. Possible modes of capturing metal ions by aminopropylphosphonate chelate
32 ligands
33
34
35
36
37
38
The strong band at 1071 cm-1 is assigned to the PO3H- group in zwitterionic CPZ 3. The
39
40 presence of peaks at 972, 968 and 958 cm-1 in the IR spectra of CPD 4, Cu-loaded and Pb-loaded
41
42 resin, respectively were due to the symmetric stretching of the PO32- motifs while the peaks for the
43
44
45 corresponding asymmetric stretching vibrations were displayed at 1062, 1046 and 1007 cm-1 (Fig.
46
47 2b-d) (Sahni et al. 1985; Cabeza et al. 2002). The strong perturbation of P-O peaks in Cu- and Pb-
48
49
50 loaded resin implies a stronger interaction amoong the metal ions and the phosphonate groups
51
52 (Sheals et al. 2001; Kolodynska et al. 2009).
53
54
55
56
57
58
59
60
61
62
63
64 12
65
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 Fig. 2 IR spectra of (a) CPZ 3, (b) CDP 4, (c) CDP 4 - Cu2+ and (d) CDP 4 loaded
48 with Pb2+
49
50
51
52 The solutions of metal nitrates were used to carry out adsorption study; as such the vibration
53
54 at 1384 cm−1, attributed to nitrate group (Sahni et al. 1985), suggests the resin’s ability to exchange
55
56
57 anion as well (Fig. 2c and d) (Kolodynska et al. 2009). The resin has 10 mol% cross-linker
58
59
60
61
62
63
64 13
65
1
2
3
4 containing pH-independent quaternary nitrogens of ≈ 20 mol% which serves the purpose of anion
5
6
7 exchange centers.
8
9 At lower loads, the chelating motifs in the resin may switch to tridentante ligands as shwn in
10
11
12 C and E or as bidentate ligands as depicted by B and D. At higher metal ion loads, ionic bonding
13
14 may lead to A (Scheme 2) (Kertman et al. 1995).
15
16
17
18
3.4 Kinetics of Adsorption
19
20
21
22 3.4.1 Order of kinetics
23
24
25 The metal removal capacity qM2+ versus temperatures plots at various time intervals are displayed
26
27
in Fig. 3a. The increase of qM2+ with time becomes asymptotic at ≈ 1h thus indicating the faster
28
29
30 attainment of the equilibrium: M2+ + Resin ⇋ Resin →M2+.
31
32
To explore the adsorption properties of the resin (Liu et al. 2010), the adsorption data were
33
34
35 fitted into Lagergren adsorption equation (Eq. 3) (Lagergren 1898) (Fig. 3b), and pseudo-second
36
37 order equation (Eq. 4) (Ho and McKay 1999) (Fig. 3c):
38
39
k1t
40 log (qe  qt )  log qe  (3)
41 2.303
42
43
t 1 t
44  2
 (4)
45 qt k2 qe qe
46
47
48 where k1 and k2 are the rate constants, while qt and qe, represent the respective adsorption capacities
49
50 of M2+ at time t and at equilibrium. Better regression values (R2), and the closeness of the qcal and
51
52
53 qexps values for both Pb2+ and Cu2+ (Table 1) support the second order kinetics (Fig. 3c) involving
54
55 a chemical adsorption process (Ramesh et al. 2007).
56
57
58 The higher rate constant (k2) and (qexp) (Table 1, Fig. 3c) for the adsorption of Cu2+ could be a
59
60 result of its much lower non-hydrated radius of 0.73 Å in comparision to 1.19 Å for Pb2+ (Lide
61
62
63
64 14
65
1
2
3
4 1998). The higher charge density of the smaller Cu2+ leads to better chelation as it enjoys the
5
6
7 entropically favorable release of water molecules from its larger hydration shell.
8
9
10
11 7.0
12 (a) (b)
13 0.5
Cu2+ 295 K; 308 K; 323 K
14

log[(qe-qt), mmol g-1]


6.0
15 Pb2+ 295 K; 308 K; ∆ 323 K
qM2+ (mmol g-1)

16
-0.5
17
5.0 Cu2+ 295 K; 308 K; 323 K
18
19
20 -1.5
21 4.0
22
23
24 3.0 -2.5
25
26 Pb2+ 295 K; 308 K; ∆ 323 K Lagergren first-order kinetics
27 2.0 -3.5
28 0 1 2 3 4 5 6 0 1 2 3 4
29
30 Time (h) Time (h)
31 7
32 (c) (d)
33 Pb2+ 295 K; 308 K; ∆ 323 K
34 1.6 Cu2+ 295 K; 308 K; 323 K 6
35
t /q t (h g mmol-1)

36
37 1.2 5
qM2+ (mmol g-1)

38
39
40 Cu2+ 323 K; 308 K; 295 K
41 0.8 4
Pb2+ 323 K; 308 K; ∆ 295 K
42
43
44 0.4 3
45
46 Pseudo second-order kinetics
47 0.0 2
48 0 2 4 6 0 0.5 1 1.5 2 2.5
49
50 Time (h) t1/2 (h-1 )
51
52
53 Fig. 3 Fig. 3 For adsorption from 0.1 M metal ion solution on to CDP 4 at pH 5.3: (a) Kinetic
54 plots of Cu2+ at ● 295, ■308, ▲323 K and Pb2+ at ○ 295, □308, ∆323 K; kinetic
55 plots of (b) First-order, (c) Pseudo second-order, (d) Intraparticle diffusion.
56
57
58
59
60
61
62
63
64 15
65
1
2
3
4 3.4.2 Intraparticle diffusion
5
6
7
8 A number of models describe the metal uptake that include the three main consecutive steps of (1)
9
10 the diffusion of metal ions through the liquid film surrounding the resin particle, (2) intraparticle
11
12
13 diffusion of metal ions through the liquid in the resin pores and (3) mass action involving metal
14
15 ions uptake onto the resin’s active sites. Intraparticle diffusion step is described by eq (5) (Weber
16
17
and Morris 1963):
18
19
20 qt  xi  kp t 0.5 (5)
21
22
23 where kp, represents the rate constant of the metal uptake, while qt denotes adsorption at verious
24
25
26 time (t) where xi is proportional to boundary layer thickness (Kavitha and Namasivayam 2007).
27
28 Fig. 3d shows the qt versus t0.5 plots with positive values of intercepts (xi ≠ 0), thereby implying
29
30
31 intraparticle diffusion as the rate-limiting step since the plots does not pass through the origin (xi
32
33 = 0) (Wu et al. 2009). The positive values of xi is a manifestation of film diffusion which implies
34
35
36
instantaneous adsorption of metal ions at an apparent t = 0. The adsorption thus takes place by
37
38 both the film and intraparticle diffusion. The plots (Fig. 3d) reveal the rapid initial uptake of metal
39
40 ions, followed by slow intraparticle diffusion and the establishment of equilibrium as described by
41
42
43 the horizontal lines.
44
45 The initial adsorption factor (Ri) is expressed by Eq. (6):
46
47
xi
48 Ri 1 (6)
49 qe
50
51
52 Table 1 reveals the larger xi values for Cu2+ than Pb2+ thereby suggesting the greater control of
53
54 film diffusion in the overall rate-limiting step for the Cu2+ ions.
55
56
57
58
59
60
61
62
63
64 16
65
1
2
3
4
Table 1 Kinetic and intraparticle diffusion parameters for the uptake of Cu2+ and Pb2+ ions
5
6 Temp qe, exp k2 ha qe, calc Ea
7 Metal ion R2
8 (K) (mmol g-1) (g mmol-1 h-1) ( mmol g-1 h-1) (mmol g-1) (kJ mol-1)
9
Pseudo second-order
10
11 Cu2+ 295 5.60 4.74 152 5.65 0.9996 31.1
12 308 5.76 11.1 370 5.79 1.0000
13
14 323 6.10 16.7 625 6.12 0.9999
15
16
Pb2+ 295 3.30 3.60 41.2 3.36 1.0000 35.5
17
18 308 3.39 6.84 80.0 3.42 0.9997
19 323 3.49 11.0 135 3.51 0.9998
20
21
22 Temp qe, exp k1 qe, calc
23 Metal ion R2
24 (K) (mmol g-1) (h-1) (mmol g-1)
25 Pseudo first-order
26
27 Cu2+ 295 5.60 2.23 1.499 0.9907
28 308 5.76 1.87 0.653 0.9479
29
30 323 6.10 4.27 0.924 0.9779
31
32 Pb2+ 295 3.30 1.81
33 1.33 0.9248
34 308 3.39 1.33 0.674 0.9925
35 323 3.49 1.89 0.609 0.9601
36
37
38 Temp kp xi qe,
39 Metal ion -1 -1/2 -1
Ri R2
40 (K) mmol g h (mmol g ) (mmol g-1)
41 Intraparticle diffusion model
42
43 Cu2+ 295 2.23 3.62 5.60 0.354 0.9816
44 308 2.01 4.40 5.76 0.236 0.9950
45 323 2.28 4.64 6.10 0.239 0.9886
46
47
48 Pb2+ 295 0.863 2.11 3.30 0.361 0.9968
49
50 308 0.789 2.49 3.39 0.265 0.9506
51 323 0.996 2.64 3.49 0.244 0.9759
52
53
54
a
Initial adsorption rate h = k2 qe2.
55
56
57
58
59
60
61
62
63
64 17
65
1
2
3
4 For the adsorption of Cu2+ at 295 K, Ri value of 0.354 indicates that the rapid film diffusion
5
6
7 controls 64.6% of the metal adsorption, while 35.4% adsorption is dictated by intraparticle
8
9 diffusion (Table 1).
10
11
12
13 3.4.3 Energy of activation Ea
14
15
16 Using second-order rate constants (k2), the ln k2 – 1/T plots (Fig. 4a) provided the respective Ea
17
18 values of 31.1 and 35.5 kJ mol-1 as per eq (7) for the uptake of Cu2+ and Pb2+ ions (Table 1). These
19
20
21 relatively smaller Ea values make the adsorption a favorable process (Veli and Pekey 2004).
22
23 Ea
24 ln k 2    constant (7)
25 2.303 RT
26
27
28
29
3.5 Adsorption isotherms
30
31
32
33 The equilibrium adsorption capacity (𝑞𝑒 ) of CDP 4 increases with increase in the initial metal ion
34
35 concentrations (𝐶0 ) (Fig.4b). Adsorption isotherm models which express 𝑞𝑒 − 𝐶𝑒 (i.e. concentration
36
37
38 at equilibrium) dependency is important in elucidation of the mechanistic pathway of the
39
40 adsorption process. Following isotherm models are considered:
41
42
Ce Ce 1
43 Langmuir:   (8)
44 qe qm qmb
45
46
1
47 Freundlich: log qe  log k f  log Ce (9)
48 n
49
50
RT RT
51 Temkin: qe  ln A  ln Ce  B ln A  b ln Ce (10)
52 b b
53
54 1 2
55 Dubinin-Radushkevich (D-R): ln qe  ln qD  BD [ RT ln( 1  )]  ln qD  BD 2 (11)
56 Ce
57
58 1
Where E 1
(12)
59
2
60 (2 BD )
61
62
63
64 18
65
1
2
3
4 The relevant linear plots of the models are shown in Fig. 4 (c-f), and the extracted parameters are
5
6
7 included in Table 2.
8
9 Langmuir isotherm is based on a monolayer formation on the homogeneous resin surface
10
11
12 where the adsorption sites are energetically similar and the adsorbed ions do not interact among
13
14 themselves (Kundu and Gupta 2006). The adsorption and desorption rate ratio is described by
15
16
17 Langmuir constants b, while Qm denotes maximum metal removal as determined from the plot of
18
19 Langmuir isotherm Eq. (8) (Fig. 4c) (Chong and Volesky 1995). A separation factor RL is
20
21 calculated using Eq. (13) (Webber and Chakkravorti 1974):
22
23
24 1
25 RL  (13)
26 (1  bC0 )
27
28 The positive RL values of < 1 indicates the favorability of the adsorptions of both metal ions (Table
29
30
31 3). An unfavorable and irreversible processes are revealed by RL ≥ 1 and RL = 0, respectively. As
32
33 noted in Table 3, RL value decreases with increasing metal ion concentration (Co); this is expected
34
35
36 since greater Co values favor adsorption. Note that the lower RL values of Cu2+ than Pb2+ confirm
37
38 the more favorable adsorption of the former ions. The resin, 1 g of which is calculated to have 3.39
39
40
41
mmol of aminopropylphosphonate ligands, has high qm value of 6.76 mmol g-1 for Cu2+ ions and
42
43 7.18 mmol for Pb2+ ions (Table 2). Each monomer unit is thus able to accommodate about two
44
45 metal ions.
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 19
65
1
2
3
4
4.00
5
(a) Arrhenius plot (b)
6 6.0
ln (k2, h-1 g mmol-1)
7
8 3.00

qM 2+ (mmol g-1)
9
10 4.0
11 2.00
12
13 2.0
14 1.00 Pb2+ Pb2+
15 Cu2+ Cu2+
16
17 0.00 0.0
18 3.00 3.20 3.40 3.60 0 0.05 0.1 0.15
19 -3
(1/T) 10 3 -1
(K ) Co (mol dm )
20
21 0.030 1.00
22 (c) Langmuir Isotherm (d) Freundlich Isotherm
23
0.80
log (qe, mmol g-1)
Ce / qe (g mL-1)

24
25 0.020
26 0.60
27
28
0.40
29 0.010
30 Pb2+
31 Cu2+ 0.20 Pb2+
32 Cu2+
33 0.000 0.00
34 0.00 0.05 0.10 -2.0 -1.8 -1.6 -1.4 -1.2 -1.0 -0.8
35
36 Ce (mol dm-3) log (ce, mol dm-3)
37
38 8.0 2.50
39 (e) Temkin Isotherm (f)
40 Dubinin – Radushkevish Isotherm
6.0 2.00
41
ln (qe, mmol g-1)

42 Pb2+
1.50
Cu2+
qe (mmol g-1)

43 4.0
44 1.00
45
2.0
46
0.50
47
48 0.0 Pb2+ 0.00
49
Cu2+
50
-2.0 -0.50
51
-4.50 -4.00 -3.50 -3.00 -2.50 -2.00 3.E+07 7.E+07 1.E+08 2.E+08
52
53
54
ln (Ce, mol dm-3) ε2
55 Fig. 4
56 Fig. 4 (a) Arrhenius plots, (b) Co versus qm of CDP 4 for Pb2+ and for Cu2+ (at pH 5.3) for
57 24 h at 295 K, (c) Langmuir, (d) Freundlich, (e) Temkin, and (f) Dubinin-
58
59 Radushkevich isotherm for the uptake of Cu2+ and Pb2+ by CDP 4 at 295 K .
60
61
62
63
64 20
65
1
2
3
4 Uniform energy heterogeneous adsorption with is described by Freundlich isotherm (Eq.
5
6
7 9); the plots provided the Freundlich constants kf and n (Fig. 4d) (Table 2).
8
9
10
11
12 Table 2 Isotherm model constants for the adsorption Cu2+and Pb2+ ions
13 at 295 K
14
15
16 Langmuir isotherm
17
Qm b
18 Metal ion R2
19 (mmol g-1) (L mol-1)
20
Pb2+ 7.18 9.16 0.9924
21
22 Cu2+ 6.76 51.0 0.9811
23
24
25
26 Freundlich isotherm
27
28 kf
Metal ion n R2
29 (mmol1-1/n g-1 L1/n)
30
31 Pb2+ 1.39 19.0 0.9990
32
33
Cu2+ 3.16 12.1 0.9829
34
35
36
37
Temkin isotherm
38 A B
39
40
Metal ion (L g-1) (kJ mol-1) R2
41 Pb2+ 111 1.37 0.9813
42
43 Cu2+ 726 1.33 0.9566
44
45 Dubinin-Radushkevich isotherm
46
47 qD E
R2
48 Metal ion mmol g-1 kJ mol-1
49
50 Pb2+ 6.49 7.29 0.9979
51
52 Cu2+ 7.29 11.4 0.9560
53
54
55
56
A favorable chemisorption process is suggested by the n values found in the range 1 – 10 (Webber
57
58 and Chakkravorti 1974).
59
60
61
62
63
64 21
65
1
2
3
4
5
6
7
8
9
10 Table 3. The RL values at various Co.
11
12 RL value
13
14 Co Pb2+ Cu2+
15 (mol dm-3)
16
17
18
0.02 0.8451 0.4949
19
20 0.04 0.7318 0.3288
21
22
23 0.06 0.6452 0.2462
24
25 0.08 0.5770 0.1967
26
27
28 0.1 0.5218 0.1638
29
30
31
32
33
34 Temkin isotherm model (Eq. 11), where R, T and A represent the gas constant (8.314 J mol-1
35
36 K-1), temperature (K) and the binding constant (L g-1), assumes that the heat of adsorption
37
38
39
decreases with increasing surface coverage by the metal ions (Liu et al. 2011). B (= RT/b) reflects
40
41 the heat of adsorption of the metal ions and are calculated from the relevant plot (Fig. 4e) (Table
42
43 2).
44
45
46 The very good fitting of the adsorption data, as reflected by R2 values, with all three isotherm
47
48 models presented suggests the formation of a monolayer on a heterogeneous surface (Table 2).
49
50
51 D-R isotherm model, represented by Eq. (11), where BD represents the free energy of
52
53 adsorption per mole of the metal ions and 𝑞𝐷 is related to the adsorption degree, is helpful in
54
55
56 estimating the resin’s porosity and the energy of adsorption E (Ali et al. 2015; Dubinin et al. 1947).
57
58 The higher values of 𝑞𝐷 , as extracted from Fig. 4f, for both metal ions match with the Qm values
59
60
61
62
63
64 22
65
1
2
3
4 as obtained from Langmuir isotherm. Both models suggest high metal uptake by the resin. The
5
6
7 favorability of the adsorption process is confirmed by the E values (Table 2) (Helfferich 1962).
8
9
10 3.6 Adsorption thermodynamics
11
12
13
14 The increase in qM2+ with a raise in temperature suggests the adsorption as an endothermic process
15
16
17 (Fig. 5a). With increasing temperature, the swelled gel provides easy access to its active sites,
18
19 thereby increasing the adsorption capacity (Cabeza et al. 2002).
20
21
22 The thermodynamic parameters Hº, Sº and Gº (= Hº - TSº) associated with the
23
24
25 adsorptions were calculated using Eq (14) and log (qe/Ce) versus 1/T plots (Fig. 5b):
26
27
qe H o S o
28 log ( )    (14)
29 Ce 2.303 RT 2.303 R
30
31
32
33 8.00 2.0
34 (a) Effect of Temperature (b) Vant-Hoff's plot
35 7.00 1.9
log(qe/Ce, L g-1)

36
qM2+ (mmol g-1)

37 6.00 1.8
38 y = -0.1511x + 2.3336
39 R² = 0.9661
5.00 Pb2+ 1.7 Pb2+
40
Cu2+ Cu2+
41
42 4.00 1.6
43
44 3.00 1.5 y = -0.0919x + 1.8838
45 R² = 1
46 2.00 1.4
47 280 300 320 340 3.00 3.20 3.40 3.60
48
49 Temperature (K) (1/T)× 103 (K-1)
50
51
52 Fig.
Fig. 55 (a) Adsorption capacity of CDP 4 – temperature plot, (b) Vant-Hoff plot.
53
54
55
56 The endothermic nature of the process is reflected by positive Hº, while negative Gº values
57
58 (Table 4) suggest the favorability of the metal uptake process. At higher temperatures, the overall
59
60
61
62
63
64 23
65
1
2
3
4 negative charges on the resin increases because of increased deprotonation of CPZ 3 to CPD 4;
5
6
7 increased interaction with the metal ions thereby shifts Gº values towards more negative side.
8
9
10
The adsorption process is accompanied by increasing randomness as described by positive Sº
11
12 because of the depletion of water molecules from metal ions’ hydration shells (Table 4).
13
14
15
16
17 Table 4 Thermodynamic data for the adsorption process
18
19
Metal Temp ΔG ° ΔH ° ΔS °
20 R2
21 ion (K) (kJ mol-1) (kJ mol-1) (J K-1mol-1)
22
23 Pb2+ 295 -8.88 1.76 36.1 0.9611
24
25 308 -9.35
26 323 -9.89
27
28 Cu2+ 295 -10.3 2.89 44.7 0.9998
29 308 -10.9
30
31 323 -11.5
32
33
34
35
36 3.7 XRD analysis
37
38
39
40 The amorphous nature of unloaded resin 4 as well as metal ions loaded resins was supported by
41
42 XRD analysis (Fig. 6) which revealed the absence of sharp peak (Tager 1978). The presence of
43
44
45 metal ions in the resin leads to a change in colour as shown in Fig. 6.
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 24
65
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
Fig. 6 XRD patterns of (a) CPD 4, (b) Pb2+-loaded and (c) Cu2+-loaded CPD 4.
27
28
29
30 3.8 SEM and EDX
31
32
33
34 CDP 4 -loaded Pb2+ and Cu2+ ions was prepared by immersing it in 0.1 M solutions of Pb(NO3)2
35
36
37
and Cu(NO3)2 at pH 5.3 for overnight. The dried samples, sputter-coated with a thin film of gold,
38
39 were scanned. EDX spectroscopy revealed the presence of several expected elements including
40
41 Na in the unloaded resin (Fig. 7a), while the Na+ ions are exchanged with Cu2+ and Pb2+ in the
42
43
44 loaded resins (Fig. 7b and c).
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64 25
65
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 Fig. 7 EDX and SEM for CDP 4: (a) unloaded, and loaded with (b) Pb2+, (c) Cu2+.
46
47
48
49
50 3.9 Treatment of industrial wastewater
51
52
53
54 CPD 4 was used for the removal of metal ions from an industrial wastewater having a pH of 7.1.
55
56
57 Wastewater (20 mL) was treated with resin 4 (50 mg) alone as well as spiked with 10 mg/L each
58
59 of Cu2+ and Pb2+ ions at room temperature for 24 h.
60
61
62
63
64 26
65
1
2
3
4
5 Table 5. Concentration of metal ions in wastewater before and
6 after adsorption with CPD 4
7
8 After treatment (g L-1)
9
10 Metal Original sample (g L-1) Original sample spiked with
11
12 Cu2+ and Pb2+
13
14
15 0 10,000 (g L-1)a
16 Al 328 7.94 10.1
17
18 Cr 12.0 7.01 7.68
19 Fe 558 573 686
20
21 Co 8.22 3.99 5.51
22 Ni 60.7 13.3 14.2
23 Cu 6.13 3.59 663
24
25 Cd <MDLb <MDL <MDL
26 Sn 6.58 1.58 0.786
27
28 Pb 0.347 0.220 796
29
a
30 Spiked samples containing 10,000 (g L-1) each of the metal
31 ions. b MDL: the method detection limit.
32
33
34
35
36 The resin treatment decreased the concentration of Al, Cr, Co, Ni, Sn considerably in the
37
38 original wastewater sample, while Cu2+ and Pb2+ concentrations in the spiked samples are
39
40
41 drastically reduced (Table 5). Note that the concentration of Al is reduced from 328 ppb to 7.94
42
43 ppb after treatment with the resin. This is remarkable since aluminum is a toxic metal which is
44
45
46
believed to lead to several clinical and neuropathological diseases, such as Alzheimer’s,
47
48 Parkinson’s, diabetes and cancer (Polizzi et al. 2000). The World Health Organization (WHO) has
49
50 has set 0.2 mg L-1 (i.e. 0.2 ppm or 200 ppb) as the maximum permissible concentration of
51
52
53 aluminum in drinking water (Veríssimo and Gomes 2008).
54
55
56
57
58
59
60
61
62
63
64 27
65
1
2
3
4
5
3.10 Recycling of theCPD 4
6
7
8
9 Three repeated cycle of adsorption/desorption processes imparted desorption efficiencies of ≥ 90%
10
11 for the metal ions (see experiment section). The resin shows stable efficiency for
12
13
14
adsorption/desorption cycles.
15
16
17
18 Table 6 A comparative adsorption capacities of CDP 4 with other types of adsorbents
19
20
21 Qm for Pb2+ Qm for Cu2+
22
Adsorbent Ref.
(mmol g-1) (mmol g-1)
23
24 Chitosan/graphene oxide nanocomposite 2.23 6.67 52
25
26
27
iminodiacetic acid / glycidyl methacrylate 0.66 2.36 19
28
29 Cystein-montmorillonite/alginate biopolymer 0.54 1.58 53
30
31
32 Layered Double Hydroxide/ MoS4 2− 2.41 2.85 54
33
34
35 N-vinyl 2-pyrrolidone/itaconic acid hydrogels 0.6 2.1 55
36
37
38 Nano-NaX Zeolite 2.23 2.28 56
39
40
41 γ-MnO2 nanostructure 0.965 1.31 57
42
43
44 Resin containing aminomethylphosphonate 2.76 2.22 20
45
46
47
48 This
Resin containing aminopropylphosphonate 7.18 6.76
49 work
50
51
52
53
54 4 CONCLUSION
55
56
57 The newly synthesized resin containing aminopropylphosphonate chelating ligands has exhibited
58
59
60 remarkable efficiencies in removing toxic metal ions. Application of the resin in aqueous medium
61
62
63
64 28
65
1
2
3
4 resulted in a very impressive performance with an adsorption capacity of 6.76 and 7.18 mmol g-1
5
6
7 for Cu2+ and Pb2+, respectively, which accorded it a prestigious place among other adsorbents
8
9 (Table 6).
10
11
12 The metal removal process followed second-order kinetics and involved rapid film and slower
13
14 intraparticle diffusion. The positive Hº and negative Gº values along with relatively smaller Ea
15
16
17 reflected the endothermic nature of the favorable adsorption process. The resin was effectively
18
19 recycled and reused. The superiority of the newly synthesized resin over several adsorbents in
20
21
22 related works is confirmed in Table 6.
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
64 29
65
1
2
3
4
5 Acknowledgments
6
7 The financial support from KFUPM project # NUS15103/4 and research facilities provided by
8
9
10 King Fahd University of Petroleum & Minerals are gratefully acknowledged.
11
12
13 References
14
15 1. Ahmad NF, Kamboh MA, Nodeh HR, Halim SNBA, Mohamad S (2017) Synthesis of
16
17 piperazine functionalized magnetic sporopollenin: a new organic-inorganic hybrid material
18
19 for the removal of lead(II) and arsenic(III) from aqueous solution, Environ Sci Pollut Res
20
21 24(27):21846–21858
22 2. Alexandratos SD, (2009) Ion-exchange resins: a retrospective from industrial and
23
24 engineering chemistry research. Ind Eng Chem Res 48(1):388–398
25
26 3. Ali SA, Ahmed SZ, Hamad Z (1996) Cyclopolymerization studies of diallyl- and
27
28 tetraallylpiperazinium salts. J Appl Polym Sci 61(7):1077−1085
29
30 4. Ali SA, Kazi IW, Ullah N (2015) New chelating ion-exchange resin synthesized via the
31
32 cyclopolymerization protocol and its uptake performance for metal ion removal. Ind Eng
33 Chem Res 54(40):9689–9698
34
35 5. Butler GB (2000) Cyclopolymerization, J Polym Sci A Polym Chem 38 (19): 3451-3461
36
37 6. Cabeza A, Ouyang X, Sharma CVK, Aranda MAG, Bruque S, Clearfield A (2002)
38
39 Complexes Formed between Nitrilotris (methylenephosphonic acid) and M2+ Transition
40
41 Metals: Isostructural Organic− Inorganic Hybrids. Inorg Chem 41(9):2325–2333
42
43
7. Chong KH, Volesky B (1995) Description of two-metal biosorption equilibria by
44 Langmuir-type models. Biotechnol Bioeng 47(4):451–460
45
46 8. Deepatana A, Valix M (2006) Recovery of Nickel and Cobalt From Organic Acid
47
48 Complexes: Adsorption Mechanisms of Metal-Organic Complexes Onto
49
50 Aminophosphonate Chelating Resin. J Hazard Mater 137(2):925-933
51
52 9. Dubinin MM, Zaverina ED, Radushkevich LV (1947) J Phy Chem 21:1351-1362.
53
54
10. Fu F, Wang Q (2011) Removal of heavy metal ions from wastewaters: a review. J
55 Environ Manage 92(3):407-418
56
57 11. Hao YM, Chen M, Hu Z (2010) Effective removal of Cu (II) ions from aqueous solution
58
59 by amino-functionalized magnetic nanoparticles. J Hazard Mater 184(1-3):392-399
60
61
62
63
64 30
65
1
2
3
4 12. He K, Chen Y, Tang Z, Hu Y (2016) Removal of heavy metal ions from aqueous solution
5
6 by zeolite synthesized from fly ash. Environ Sci Pollut Res 23(3):2778–2788
7
8 13. Helfferich F (1962) Ion Exchange. McGraw-Hill, New York.
9
10 14. Ho YS, McKay G (1999) Pseudo-second order model for sorption processes. Process
11
12 Biochem 34(5):451–465
13
14
15. Hua M, Zhang S, Pan B, Zhang W, Lv L, Zhang Q (2012) Heavy metal removal from
15 water/wastewater by nanosized metal oxides: a review. J Hazard Mater 211–212:317-331
16
17 16. Jamiu ZA, Saleh TA, Ali SA (2015) Synthesis of a unique cross-linked
18
19 polyzwitterion/anion with an aspartic acid residue and its use for Pb2+ removal from
20
21 aqueous solution. RSC Adv 5(53):42222-42232
22
23 17. Jiang J, Ma XS, Xu LY, Wang LH, Liu GY, Xu QF, Lu JM, Zhang Y (2015)
24
25
Applications of chelating resin for heavy metal removal from wastewater. e-Polym.
26 15(3):161-168
27
28 18. Kavitha D, Namasivayam C (2007) Experimental and kinetic studies on methylene blue
29
30 adsorption by coir pith carbon. Bioresour Technol 98(1):14–21
31
32 19. Kayal N, Singh N (2007) Stepwise complexometric determination of aluminium, titanium
33
34 and iron concentrations in silica sand and allied materials. Chem Cent J 1:24-28
35
20. Kertman SV, Kertman GM, Leykin YA (1995) A thermochemical study of complex
36
37 formation in chelating ion-exchange resins. Thermochim Acta 256 (2):227-235
38
39 21. Kołoynska D, Hubicki Z, Geca M (2008) Application of a new generation complexing
40
41 agent in removal of heavy metal ions from aqueous solutions. Ind Eng Chem Res
42
43 47(9):3192-3199
44
45 22. Kolodynska D, Hubicki Z, Pasieczna-Patkowska S (2009) FT-IR/PAS studies of Cu (II)–
46 EDTA complexes sorption on the chelating ion exchangers. Acta Phys Pol A 116(3):340–
47
48 343
49
50 23. Kudaibergenov S, Laschewsky WJ, Laschewsky A (2006) Polymeric Betaines: Synthesis,
51
52 Characterization, and Application. Adv Polym Sci 201:157-224.
53
54 24. Kundu S, Gupta AK (2006) Arsenic adsorption onto iron oxide-coated cement (IOCC):
55
56 regression analysis of equilibrium data with several isotherm models and their
57 optimization. Chem Eng J 122(1-2):93–106.
58
59 25. Lagergren S, Sven K (1898) Vetenskapsakad Handl 24:1–39.
60
61
62
63
64 31
65
1
2
3
4 26. Lide DR (1998) Handbook of Chemistry and Physics, 79th ed, CRC Press, Boca Raton.
5
6 27. Liu J, Ma Y, Xu T, Shao G (2010) Preparation of zwitterionic hybrid polymer and its
7
8 application for the removal of heavy metal ions from water. J Hazard Mater 178(1-3):1021-
9
10 1029
11
12 28. Liu J, Song L, Shao G (2011) Novel zwitterionic inorganic-organic hybrids: Kinetic and
13
14
equilibrium model studies on Pb2+ removal from aqueous solution. J Chem Eng Data
15 56(5):2119-2127.
16
17 29. Lofrano G, Carotenuto M, Libralato G, Domingos RF, Markus A, Dini L, Gautam RK,
18
19 Baldantoni, D, Rossi M, Sharma SK, Chattopadhyaya MC, Giugni M, Meric S (2016)
20
21 Polymer functionalized nanocomposites for metals removal from water and wastewater:
22
23 An overview. Water Res. 92:22-37.
24
25
30. Marzougui Z, Chaabouni A, Elleuch B, Elaissari A (2016) Removal of bisphenol A and
26 some heavy metal ions by polydivinylbenzene magnetic latex particles. Environ Sci Pollut
27
28 Res 23(16):15807–15819.
29
30 31. McGrew FC (1958) Structure of synthetic high polymers. J Chem Educ 35(4):178-186.
31
32 32. Ministry of Health of P. R. China. Standards for drinking water quality. (2006) GB 5749.
33
34 33. Mishra SP (2014) Adsorption-desorption of heavy metal ions Curr Sci 107(4):601-612.
35
34. Ngah WSW, Hanafiah MA (2008) Removal of heavy metal ions from wastewater by
36
37 chemically modified plant wastes as adsorbents: a review. Bioresour Technol
38
39 99(10):3935–3948.
40
41 35. Polizzi S, Pira E, Ferrara M, Bugiani M, Papaleo A, Albrera R, Palmi S (2000) Neurotoxic
42 effects of aluminium among foundry workers and Alzheimer's disease. Neurotoxicology
43
44 23(6):761-774.
45
46 36. Ramesh A, Hasegawa H, Maki T, Ueda K (2007) Adsorption of inorganic and organic
47
48 arsenic from aqueous solutions by polymeric Al/Fe modified montmorillonite. Sep Purif
49
50 Technol 56(1):90−100.
51
52
37. Ranade V, Bhandari V (2014) Industrial Wastewater Treatment, Recycling and Reuse,
53 Butterworth-Heinemann, ISBN 13: 9780080999685.
54
55 38. Rao Z, Feng K, Tang B, Wu P, (2017) Surface Decoration of Amino-Functionalized Metal–
56
57 Organic Framework/Graphene Oxide Composite onto Polydopamine-Coated Membrane
58
59
60
61
62
63
64 32
65
1
2
3
4 Substrate for Highly Efficient Heavy Metal Removal. ACS Appl Mater Interfaces
5
6 9(3):2594–2605.
7
8 39. Ray PZ, Shipley HJ (2015) Inorganic nano-adsorbents for the removal of heavy metals and
9
10 arsenic: a review. RSC Ad. 5(38):29885-29907.
11
12 40. Sahni SK, Bennekom RV, Reedijk J (1985) A spectral study of transition-metal complexes
13
14
on a chelating ion-exchange resin containing aminophosphonic acid groups. Polyhedron
15 4(9):1643–1658.
16
17 41. Samiey B, Cheng C-H, Wu J (2014) Organic-inorganic hybrid polymers as adsorbents for
18
19 removal of heavy metal ions from solutions: A review. Materials 7(2):673-726.
20
21 42. Sheals J, Persson P, Hedman B, (2001) IR and EXAFS Spectroscopic studies of glyphosate
22
23 protonation and copper(II) complexes of glyphosate in aqueous solution. Inorg Chem,
24
25
40(17):4302-4309
26 43. Shen X, Qiu G, Yue C, Guo M, Zhang M (2017) Multiple copper adsorption and
27
28 regeneration by zeolite 4A synthesized from bauxite tailings. Environ Sci Pollut Res
29
30 24(27):21829–21835
31
32 44. Singh PK, Singh VK, Singh M (2007) Zwitterionic polyelectrolytes: A Review. e-Polym
33
34 30:1–34
35
45. Tager A (1978) Physical Chemistry of Polymer. Mir Publisher, Moscow.
36
37 46. US Environmental Protection Agency. National primary drinking water regulations. (2009)
38
39 (EPA 816-F-09-004).
40
41 47. Veli S, Pekey B (2004) Removal of copper from aqueous solution by ion exchange
42
43 resins. Fresenius Environ Bull 13(3b):244–250.
44
45 48. Veríssimo MIS, Gomes MTSR (2008) The quality of our drinking water: aluminium
46
47 determination with an acoustic wave sensor. Anal Chim Acta 617(1-2):162-166.
48
49 49. Wang Q, Gao W, Liu Y, Yuan J, Xu Z, Zeng Q, Li Y (2014) Schröder, M., Simultaneous
50
51 adsorption of Cu(II) and SO42− ions by a novel silica gel functionalized with a ditopic
52
53 zwitterionic Schiff base ligand. Chem Eng J 250:55-65.
54
55 50. Webber TW, Chakkravorti RK (1974) Pore and solid diffusion models for fixed‐bed
56
adsorbers. AlChE J 20 (2):228–238.
57
58 51. Weber WJ, Morris JC (1963) Kinetics of adsorption on carbon from solution. Eng Div Am
59
60 Soc Civ Eng 89(2):31–60.
61
62
63
64 33
65
1
2
3
4 52. Wu FC, Tseng RL, Juang RS (2009) Initial behavior of intraparticle diffusion model used
5
6 in the description of adsorption kinetics. Chem Eng J 153(1-3):1–8.
7
8 53. Xie Y, Huang Q, Liu M, Wang K, Wan Q, Deng F, Lu X, Zhang L, Wei Y (2015) Mussel
9
10 inspired functionalization of carbon nanotubes for heavy metal ion removal. RSC Adv
11
12 5(84):68430-68438.
13
14
54. Yusuf M, Elfghi FM, Zaidi SA, Abdullah EC, Khan MA (2015) Applications of graphene and
15 its derivatives as an adsorbent for heavy metal and dye removal: a systematic and
16
17 comprehensive overview. RSC Adv 5(62):50392-50420.
18
19 55. Yuwei C, Jianlong W (2011) Preparation and characterization of magnetic chitosan
20
21 nanoparticles and its application for Cu (II) removal. Chem Eng J 168(1):286-292.
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
64 34
65

You might also like