Sei sulla pagina 1di 50

Accepted Manuscript

Title: Development of cysteine amide reduced graphene oxide


(CARGO) nano-adsorbent for enhanced uranyl ion adsorption
from the aqueous medium

Authors: Swati Verma, Raj Kumar Dutta

PII: S2213-3437(17)30434-7
DOI: http://dx.doi.org/10.1016/j.jece.2017.08.047
Reference: JECE 1842

To appear in:

Received date: 27-4-2017


Revised date: 26-8-2017
Accepted date: 28-8-2017

Please cite this article as: Swati Verma, Raj Kumar Dutta, Development of cysteine
amide reduced graphene oxide (CARGO) nano-adsorbent for enhanced uranyl
ion adsorption from the aqueous medium, Journal of Environmental Chemical
Engineeringhttp://dx.doi.org/10.1016/j.jece.2017.08.047

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. 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.
Development of cysteine amide reduced graphene oxide (CARGO) nano-adsorbent for

enhanced uranyl ion adsorption from the aqueous medium

Swati Verma, Raj Kumar Dutta*

Department of Chemistry, Indian Institute of Technology Roorkee, Roorkee – 247667, India

*Corresponding Author: duttafcy@iitr.ac.in (R.K. Dutta), Tel: +91 1332 285280

GRAPHICAL ABSTRACT

UO22+
UO22+ UO22+

UO22+

pH 5 UO22+ UO22+ UO22+

UO22+ UO22+ UO22+

CARGO before adsorption CARGO after adsorption

Carbon Oxygen Nitrogen Sulfur

1
HIGHLIGHTS

 Synthesis and characterization of L-cysteine amide reduced graphene oxide

(CARGO) functionalized with multiple amines/amides as nano-adsorbent

 Uranyl ion adsorption followed pseudo second order kinetics and chemisorption

mechanism, with qmax = 338 mg/g in 60 min at pH 5

 Multiple amine/amide binding sites in CARGO, increase in BET surface area and

favorable electrostatic interaction with uranyl ions favored enhanced adsorption

 Mild interference (<10 % of qe) by Ca2+, Mg2+, HCO3- and CO32- in uranyl ion

adsorption

 The regeneration and re-usability of the CARGO adsorbent has been demonstrated

without affecting adsorption capacity

Abstract

The affinity of uranyl ions for binding with nitrogen bearing ligands has led to development

of several amine/amide reduced graphene oxides as adsorbent for uranyl ion removal. We

have developed a novel nano-adsorbent (referred as CARGO) with an aim of increasing

nitrogen groups in reduced graphene oxide (GO) for enhanced uranyl ion removal. There are

two important stages in this work: (a) synthesis of CARGO, which comprises of GO

reduction by the as synthesized L-cysteine amide and (b) optimization of uranyl ion

adsorption by CARGO. The synthesis of L-cysteine amide (precursor) is confirmed by FT-IR

and 1H NMR. Reduction of GO by L-cysteine amide is revealed from decrease in the acidic

functional groups of GO and has been characterized by XRD, UV-vis, Raman and FT-IR

spectroscopy techniques. The morphology of CARGO represented flake like structure,

revealed from FE-SEM and TEM images. The uranyl ion adsorption on CARGO surface was

confirmed from EDX analysis and supported by elemental map of uranium and carbon. The

regeneration, re-usability and selectivity of uranium adsorption by CARGO have been


2
studied. The uranyl adsorption was kinetically and thermodynamically favorable, which

accounted for 88.9 % uranyl ion removal in 60 min at optimized pH 5 via chemisorption

mechanism and the corresponding qmax value derived from Langmuir adsorption isotherm

was 337.93 mg/g. Such high uranyl ion adsorption capacity is attributed to more amine and

/or amide binding sites on CARGO, high BET surface area (57.086 m2/g) and negative zeta

potential values of the adsorbent.

Keywords:

Nano-adsorbent; Uranium; Cysteine amide; Reduced graphene oxide; Adsorption; Arsenazo-

III

1. Introduction

Uranium is radioactive and environmentally hazardous element [1]. It is used as a fuel

material in nuclear reactors. Nuclear industries are thus point sources for potential release of

uranium into the environment. In addition, uranium contents in the ground water of certain

parts of the world are elevated due to local geological and geochemical features. There has

been continued scientific interest towards developing technologies for decontaminating

uranium from water [2-5]. In this regard, uranium removal via adsorption has received most

attention owing to its simplicity and cost effectiveness of the process. A wide range of

materials have been studied as adsorbents for the removal of uranium from aqueous solution.

The adsorption capacities (qe) of some of the commonly studied adsorbents like sepiolite,

activated bentonite, activated carbon, activated charcoal, silica and zeolites ranged between 8

mg/g and 36 mg/g [6-11]. The efficiency of adsorption depends on various factors, e.g.,

surface area, pore size, functional groups and residual charge on the adsorbent. It also

depends on the size of the solute, which impacts on the slow rate of adsorption of uranium

species.

3
The uranium adsorption efficiencies were improved for nanoscale materials as

adsorbents, e.g., multi-walled carbon nanotubes (39.1 mg/g), magnetic Fe3O4@SiO2 (52

mg/g), titanium dioxide (60 mg/g), and tin oxide (66.67 mg/g) [12-15]. Increased surface area

and modification of the surface of nanomaterials with suitable capping agents improved

adsorption of uranium [16-19]. Under this category of adsorbents, graphene oxide (GO) and

reduced graphene oxide (rGO) exhibited strikingly high adsorption capacities for uranium

adsorption [20,21]. This is attributed to functionalization of graphene oxide by suitable

ligands together with increase in the surface area to volume ratio of the adsorbent for

enhanced metal ion selectivity. The structure of GO contains large number of negatively

charged oxygen bearing functional groups (e.g., phenolic group, ether and carboxylic acid

group) in the basal plane and at the edges which serves as the binding sites for the removal of

positively charged divalent metal ions from aqueous solution [22]. Since carboxylic acid

group is non-selective to uranyl ions, it would severely impact on the efficiency of uranium

removal due to interfering cations.

It is found that graphene oxide modified with amines, amide and amidoxime ligands

enhanced the uranyl ion adsorption capacities due to increase in the chemical affinity of

uranyl ions for these functional groups [23-27]. Similarly, GO composites modified with

polypyrrole and polydopamine revealed high uranium removal efficiencies, with large qe

values of 147.1 mg/g and 145.39 mg/g, respectively [28,29]. Similarly, nanocomposites of

GO with polyaniline and chitosan revealed enhanced adsorption capacities, where the qmax

values were derived from the Langmuir plot as 245.14 mg/g at pH 3 and 225.78 mg/g at pH

4, respectively [30,31]. It may be commented that enhancement in the uranyl ion adsorption

could be achieved by activating graphene based system as nano-adsorbent, especially by

incorporating multiple nitrogen bearing functional groups.

Instead of using commercially available ligands with higher number of nitrogen bear

groups, we have developed a novel multifunctional amine-amide ligand for reducing GO.

4
Here more nitrogen bearing functional groups are incorporated in the GO basal plane with an

aim of enhancing uranium adsorption capacity. The terminal carboxylic acid groups of GO

were reduced by modified amino acid, e.g., L-cysteine amide, which has been synthesized by

amidation of L-cysteine via acyl chloride. The adsorbent, referred to as cysteine amide

reduced graphene oxide (CARGO), has been thoroughly characterized and studied for uranyl

ion adsorption in the presence of common cations and anions that are usually present in

drinking water. The uranyl ion adsorption mechanism has been studied using adsorption

kinetic models and adsorption isotherms.

2. Materials and method

2.1 Materials

Graphite powder (mesh size < 20 μm) and Arsenazo (III) were procured from Sigma

Aldrich, GmbH, Germany. Concentrated sulfuric acid (H2SO4, 98%), ortho-phosphoric acid

(H3PO4, 88%), ethanol (C2H5OH, 99%), ethylene glycol (C2H4O2, 98%), diethyl ether

(C4H10O), potassium permanganate (KMnO4), sodium hydroxide (NaOH), thionyl chloride

(SOCl2), liq. ammonia (NH4OH) and L-cysteine (C3H7NO2S) were obtained from Himedia,

Mumbai, India. The uranyl nitrate hexahydrate, i.e. UO2(NO3)2.6H2O (molecular weight =

502 g) was purchased from Merck, India. All reagents used for the synthesis were of

analytical grade and were used without any further purification. De-ionized water (DI water,

Millipore) has been used throughout this study.

2.2 Synthesis of CARGO as adsorbent

Synthesis of CARGO is a multi-step process consisting of the following major steps:

Step-1: Synthesis of GO nanosheets

Firstly, nanosheets of graphene oxide (GO) were synthesized according to improved

Hummer’s method [32]. Briefly, 0.5 g graphite powder was dispersed in the pre-cooled

5
mixture of H2SO4 and H3PO4 in 9:1 volume ratio followed by the addition of 3 g KMnO4

(solid). The reaction mixture was kept for stirring with mild heating at 50-60 °C for 12 h in

water bath. This solution was then cooled to room temperature and was poured into 70 mL

ice cooled water containing 1 mL of 30 % H2O2. A solid yellowish-brown precipitate was

obtained which was separated by centrifugation and subsequently washed with 30 mL DI

water, 30 mL of 30 % v/v HCl, 30 mL ethanol and finally coagulated using diethyl ether. The

as-obtained GO nanosheets were exfoliated in DI water using ultra-sonication and left for

drying at room temperature over Petri-dish.

Graphene oxide nanosheets comprise of carboxylic acids, phenolic and lactonic

groups which contribute towards total acidic functional groups present on GO nanosheets

[33]. It is therefore important to characterize the oxygen functional groups in the as-

synthesized GO, which was quantified using Boehm titration method [34]. Briefly, 0.1 g of

GO nanosheets was dispersed in 100 mL solution 0.1 M of NaOH for the determination of

surface functional groups. The obtained dispersion was mechanically stirred for 48 h at room

temperature to reach equilibrium and then filtered to separate GO nanosheets from solution.

A 5 mL filtrate was then titrated with 0.05 M HCl to neutralized excess of NaOH using

phenolphthalein indicator. A control sample without GO nanosheets was also titrated in the

same way. The concentration of acidic functional groups in GO was found to be 0.171 mol/g.

On comparing our result with that in literature, it may be remarked that the surface of the as-

synthesized GO nanosheets was heavily decorated with oxygen containing functional groups

during the oxidation process.

Step 2: Synthesis of L-cysteine amide

The amide derivative of L-cysteine was prepared by reacting L-cysteine with thionyl

chloride and liq. NH3. Briefly, 100 mg of L-cysteine was added to 20 mL of pre-ice cooled

thionyl chloride and the mixture was refluxed at 50 °C for 30 minutes. After refluxing, the

6
contents were slowly poured over excess of ice cooled liq. NH3. The solid material obtained

was separated by filtration and re-crystallized using 1:1 water/ethanol mixture.

The synthesis of the amide derivative of L-cysteine has been confirmed by FT-IR

spectroscopy and 1H NMR studies. L-cysteine was used as a precursor for synthesizing L-

cysteine amide. The FT-IR spectrum of L-cysteine is given in Fig. S1a. The structure of L-

cysteine is given in the inset of Fig. S1a. The FT-IR spectrum of L-cysteine revealed peaks at

3445 cm-1 and 3178 cm-1 (–NH2 and –OH group), 2977 cm-1 (aliphatic C-H stretching), 2549

cm-1 (S-H stretching), 1588 cm-1 (asymmetric C=O stretching and asymmetric N-H bending),

1537 cm-1 (symmetric N-H bending) and 1393 cm-1 (aliphatic C-H bending and C-O

symmetric stretching), which are typically due to L-cysteine. The multiple bands in the

region 1500 cm-1 to 400 cm-1 corresponded to the characteristic fingerprint region of L-

cysteine molecule, which consists of bending and out of plane vibration modes within the

molecule [35]. In the case of L-cysteine amide, the FT-IR spectrum was expectedly similar to

that of L-cysteine, except for a new peak emerged at 1624 cm-1 (Fig. S1b). This peak is

attributable to C=O stretching vibration mode of amide groups. Further, the intensity of the

1588 cm-1 peak in L-cysteine amide due to carboxylic acid/carboxylate decreased sharply

owing to the reduction of carboxylic acid groups. These two observations strongly supported

the modification of L-cysteine to L-cysteine amide.

Furthermore, the reduction of L-cysteine to L-cysteine amide was addressed by 1H

NMR studies (Fig. S2). The 1H NMR spectra of L-cysteine revealed three doublet signals of

doublet signals at δ = 3.85 ppm for methine proton and δ = 2.86 ppm and 3.00 ppm for

methylene protons, which was similar to the reported literature [36]. In the case of amide

derivative, these peaks are modified and a new triplet was observed at δ = 8.30-8.55 ppm.

This proton signal is attributable to the formation of amide derivative [37].

Step 3: Reduction of graphene oxide with L-cysteine amide

7
A batch of 0.1 g freshly synthesized L-cysteine amide was mixed in 50 mL of 0.1 M

NaOH solution and was added drop-wise to 100 mL GO dispersion of concentration 1

mg/mL in stirring condition. The reaction mixture was subjected to microwave heating for 20

min at 800 W. Black colored slurry was produced and was separated by centrifugation. It was

washed repeatedly with de-ionized water till the pH of the supernatant became neutral. In this

CARGO batch, the GO: L-cysteine amide weight ratio was 1:1, and referred to as CARGO-1.

Similarly, the CARGO-2 and CARGO-3 batches were synthesized by treating 0.5 g and 1.0 g

of L-cysteine amide with 100 mL of GO dispersion so that weight ratio of GO:L-cysteine

amide were 1:5 and 1:10, respectively.

The as synthesized batches of CARGO adsorbent were thoroughly characterized by

XRD, UV-visible spectroscopy, Raman Spectroscopy, FT-IR spectroscopy and the

morphology was imaged by transmission electron microscopy. The detailed information

about the instrumentation techniques and sample preparation are provided as Supporting

Information (Section S1).

2.3. Methodology for uranium adsorption studies

Adsorption experiments were carried out in batch mode in a temperature controlled

water bath shaker with a speed of 200 rpm. The effect of adsorbent doses was studied by

treating varying amounts (i.e. 10-100 mg) of CARGO-1, CARGO-2 and CARGO-3 with 100

mL solution of uranyl ion of concentration 100 mg/L. The corresponding adsorption

efficiency of uranyl ions on the adsorbent was determined as:


𝐶𝑜 −𝐶𝑒
𝐴𝑑𝑠𝑜𝑟𝑝𝑡𝑖𝑜𝑛 𝑒𝑓𝑓𝑖𝑐𝑖𝑒𝑛𝑐𝑦 (%) = × 100 (A1)
𝐶𝑜

where, Co and Ce (mg/L) are the initial and equilibrium concentrations of uranyl ions in the

reaction medium. The adsorption capacity at equilibrium condition (i.e. qe) was determined

as:
𝐶0 −𝐶𝑒
𝑞𝑒 = ×𝑉 (A2)
𝑚

8
where, Co and Ce are same as in A1, m is the amount of adsorbent (in mg) and V is the volume

of the uranyl ions (in mL). Maximum uranium adsorption efficiency was recorded for

CARGO-1 for an optimized adsorbent dose of 50 mg in 100 mL of uranyl ions solution.

Therefore, all further adsorption studies were performed with 50 mg of CARGO-1. The effect

of parameters like pH (2-9), temperature (288-313 K) and contact time (2-120 min) on the

adsorption were optimized using 50 mg of CARGO-1 in 100 mL solution of 100 mg/L of

uranyl ions. Adsorption isotherms were obtained by treating 50 mg of CARGO-1 with 100

mL solution of uranyl ions of initial concentrations in the range 10-250 mg/L at pH 5 and at

298 K.

Uranyl ions adsorption was also studied in presence of common cations and anions

available in drinking water. The interfering ions considered for this study are Ca2+ (75 mg/L),

Mg2+ (30 mg/L), K+ (50 mg/L), Na+ (200 mg/L), Pb2+ (0.1 mg/L), Fe2+ (0.3 mg/L), Fe3+ (0.3

mg/L), V5+ (0.1 mg/L), Zn2+ (5 mg/L), CO32- (300 mg/L), HCO3- (300 mg/L), Cl- (250 mg/L),

SO42- (200 mg/L), PO43- (0.1 mg/L) and humic acid (3 mg/L) [38].

2.4. Uranium analysis by Arsenazo (III) based colorimetric method

A 1000 mg/L stock solution of uranium was prepared by dissolving 2.11 g of

UO2(NO3)2.6H2O in 5 % HNO3 prepared in de-ionized water. Desired concentrations of test

uranium solution were prepared by serial dilution with de-ionized water. The pH of the test

solutions were adjusted using 0.1 M HNO3 and NaOH. The concentration of uranyl ions in

the solution was measured by colorimetric method using Arsenazo (III) dye as a complexing

agent [39]. Briefly, a solution of arsenazo (III) reagent of concentration 0.07% was prepared

in 3M perchloric acid (70 %) and treated with an aliquot of supernatant containing uranyl

ions in 4:1 volume ratio. The absorbance of the arsenazo (III)–UO22+ complex thus formed

was measured at 650 nm by Shimadzu UV-1800 spectrophotometer which would correspond

to the concentration of free uranyl ions in a reaction medium. The peak intensity at 650 nm

9
increases with increase in the uranyl ion concentration (given as Supporting Information, Fig.

S3a) and the validity of the uranium measurement by Arsenazo (III) method was confirmed

from the linear calibration plot for uranyl concentration ranging between 1 – 25 mg/L (given

as Supporting Information, Fig. S3b).

3. Results and discussion

3.1. Characterization of CARGO adsorbent

The modification of graphene oxide by reduction with L-cysteine amide has been

studied by an array of characterization techniques. The UV-visible absorption spectrum of

graphene oxide revealed an intense absorption peak at 230 nm, corresponding to π-π*

transition associated with aromatic C=C bond (given as supporting information, Fig. S4). A

red shift for the π-π* transition peak was observed for CARGO-1, CARGO-2 and CARGO-3

at 236, 245 and 256 nm, respectively (given as supporting information, Fig. S4). The gradual

red shift occurred with increasing concentration of L-cysteine amide, which can be attributed

to the conjugated π electron system due to enhanced reduction in graphene oxide structure.

The modification of graphene oxide was also evident from the disappearance of the weak

shoulder peak at 300 nm owing to n-π* transition of peripheral carboxylic acid groups [40].

Furthermore, the weak shoulder peak which appeared for graphene oxide was not observed

for the CARGO batches. This implied that the n-π* transitions was suppressed and the π bond

character of the carboxylic acid group is likely to be modified by L-cysteine amide. Similar

phenomenon has also been reported in literature for graphene oxide reduced by different

methods [41,42]. The reduction of GO by L-cysteine amide is also evident from XRD

measurements. Firstly, the synthesis of graphene oxide from graphite is reflected from the X-

ray diffraction pattern, which revealed an intense diffraction peak at 2θ = 9.58°

corresponding to the (0 0 2) plane, owing to a well crystalline structure with inter-planar

distance of 9.22 Å (Supporting Information, Fig. S5). The d-spacing in the GO was larger

10
than the d-spacing of the (0 0 2) plane of graphite (3.34 Å). The increase in the d-spacing in

GO was attributable to introduction of oxygen containing moieties and water molecules

between graphene layers [43]. Further, the modification of GO by L-cysteine amide resulted

in the shift of the X-ray diffraction peak of the (0 0 2) plane towards higher 2θ value, i.e., at

12.04° for CARGO-1, 24.77° for CARGO-2, and 27.33° for CARGO-3. The corresponding

d-spacings of (0 0 2) plane were determined as 7.34 Å, 3.58 Å and 3.26 Å, respectively. The

decrease in the d-spacing in the CARGO-1, 2 and 3 batches confirmed reduction of graphene

oxide. The decrease in the d-spacing is more evident for the CARGO-2 and CARGO-3

batches, which were similar to that of graphite indicating a possible re-stacking of the

graphene sheets due to reduction of graphene oxide [44]. It is evident from the XRD plot that

the reduction of GO with L-cysteine amide resulted into broadening of the peak

corresponding to the (0 0 2) plane, suggested that the reduction process resulted in poorly

crystalline reduced graphene oxide nanosheets. Similar results have been reported for the

reduction of graphene oxide by L-ascorbic acid, hydrazine and sodium borohydride

[42,45,46].

The reduction of GO by L-cysteine amide was further supported from FT-IR

spectroscopic studies. The Fig. 1 corresponded to the FT-IR spectrum of GO revealed broad

peak at 3420 cm-1 (O-H stretching), weak peak at 1736 cm-1 (C=O stretching), 1637 cm-1

(conjugated C=C bending), 1222 cm-1 (phenolic C-O stretching) and 1050 cm-1 (epoxy C-O-C

stretching vibrations), which are consistent with literature reports [47]. The FT-IR spectrum

of CARGO-1 revealed broad peak at 3438 cm-1, 3140 cm-1, 3023 cm-1 and 2584 cm-1 as

reflected in Fig. 1, which were similar to those recorded for the stretching vibrations of N-H,

O-H, C-H, and S-H groups of cysteine amide (Fig. S1b). In addition, the peaks at 1584 cm-1

and 1623 cm-1 were due to the N-H bending vibrations and C=O stretching vibrations of the

amide group. The peaks at 1486 cm-1 and 1404 cm-1 corresponding to aromatic C=C and C-H

bending vibrations. Similar pattern of peaks was also observed for CARGO-2 and CARGO-3.

11
Our results are similar to the FT-IR spectra reported for various types of amide modified GO

[25,26,48,49]. It may therefore be concluded that GO nanosheets were successfully modified

by reducing with L-cysteine amide.

The reduction of GO is associated with reduction of carboxylic acid. In this regard,

the concentration of acidic functional groups in GO, CARGO-1, CARGO-2 and CARGO-3

were determined from Boehm titration as 0.171, 0.151, 0.147 and 0.146 mol/g, respectively.

There was nearly 20 % decrease in the concentration of total oxygen functional group with

respect to GO, which is attributable to the modification of the carboxylic acid of GO by the

L-cysteine amide. The decrease in the carboxylic acid in CARGO batches with respect to GO

was an indicator of reduction of GO by L-cysteine amide [50]. A likely scheme of the

synthesis of CARGO is given in Scheme 1.

The reduction of GO by L-cysteine amide is further supported by Raman

spectroscopy. The characteristic D and the G bands in GO corresponding to the sp3 and sp2

hybridized carbon atoms were observed at 1363cm-1 and 1604 cm-1, respectively (Fig. 2)

[51]. The D-bands for CARGO batches were shifted to 1357 cm-1 and lesser. These Raman

peaks corresponded to the first order scattering of the E2g mode [52]. The values of ID/IG ratio

were more for the CARGO batches than the GO. This was attributable to reduction of GO by

L-cysteine amide and our results are consistent with related literature report [41]. The

increase in the ID/IG ratio corresponded to the enhanced degree of exfoliation or deformation

in the structure of the aromatic graphene layer.

The FE-SEM images of the batches of CARGO revealed well exfoliated and wrinkled

graphene sheets (Fig. S6). A more informative morphology of CARGO adsorbent was

obtained from transmission electron microscopy (TEM) image of CARGO-1, which revealed

the formation of well exfoliated two dimensional transparent nanosheets with folded edges

(Fig. 3).

12
3.2. Adsorption of UO22+ by CARGO-1

The adsorption of uranyl ions on GO and CARGO-1 were compared under similar

conditions by taking an arbitrary amount of adsorbent and the respective kinetic data are

given in Fig. 4. The adsorption capacity at equilibrium condition (qe) was determined as

123.57 mg/g for GO and 175.13 mg/g CARGO. A significant improvement in the adsorption

capacity of GO was evident when it is modified by reducing with L-cysteine amide. Further

uranyl ion adsorption studies were conducted on CARGO batches.

3.2.1. Adsorbent dose and pH optimization

The effect of adsorbent dose of CARGO-1, CARGO-2 and CARGO-3 on the

adsorption efficiency and adsorption capacity for uranyl ions were studied. The adsorption

efficiencies were ranked as: CARGO-1 > CARGO-2 > CARGO-3 (Fig. 5a). The same pattern

was observed for BET surface area, i.e., CARGO-1 (57.088 m2/g) > CARGO-2 (51.969 m2/g)

> CARGO-3 (25.146 m2/g). It may be surmised that surface area of CARGO batches was an

important parameter for enhancing uranyl ion adsorption. For an arbitrary uranium

concentration of 100 mg/L, the adsorption efficiency of CARGO-1 was increased from 51.3

% as the adsorbent dose was increased from 10 mg to 50 mg per 100 mL. An equilibrium

condition was achieved when the adsorption efficiency was 88.9%. The increase in the

adsorption efficiency of uranyl ions with increasing concentration of adsorbent is attributed to

the availability of more number of binding sites. The corresponding adsorption capacities at

equilibrium (qe) were calculated to be in the range between 514 mg/g to 176 mg/g (Fig. 5b).

Therefore, 50 mg of CARGO-1 was the optimum adsorbent dose which corresponded to 88.9

% uranyl ion removal with a high qe value of 176 mg/g. All subsequent uranyl ion adsorption

studies were performed using the optimized dose.

Next, the effect of pH of the uranium solution was studied with respect to the adsorption

efficiency. The pH of the medium not only influences the surface charge of the adsorbent, but

13
it may also alter the chemical species of the adsorbate. It is particularly relevant for uranium

as uranyl ion (UO22+) is a dominant uranium species at pH < 5. While at higher pH, uranium

exists as hydroxy and carbonate species, such as UO2(OH)3-, UO2(OH)42-, UO2(CO3)22- and

UO2(CO3) [53]. The adsorption studies of uranyl ions on CARGO-1 were performed over a

wide pH range between 2 and 9. The maximum uranium adsorption efficiency was 89 % at

pH 5, which corresponded to adsorption capacity (qe) of 176 mg/g (Fig. 5c). The uranium

adsorption was determined from the difference in the residual uranyl concentration after

adsorption from the initial concentration of uranium. The residual uranyl concentrations for

different batches were determined from the UV-visible spectra of Arsenazo (III)-uranyl ion

complexes, as given in supporting information in Fig. S7. From Fig. 5c, it was noted that the

qe values were more than 100 mg/g over the entire pH range. It may be remarked here that the

optimized qe value for CARGO-1 was significantly improved as compared to different

graphene oxide nanostructure reported in literature [54]. This could be attributed to negative

surface charge on CARGO-1 over the entire pH range as reflected from the measured

negative zeta potentials (Fig. 5d). This favored electrostatic interaction between the uranyl

ions and the CARGO adsorbent.

3.2.2 Selectivity of the adsorbent

The effect of interfering ions is an important parameter in the study of metal ion

adsorption at solid-liquid interface. The selectivity of CARGO-1 for adsorption of uranyl ions

was examined in presence of cations and anions which are commonly present in drinking

water. The common interfering ions considered in this study were Ca2+, Mg2+, K+, Na+, Pb2+,

Fe2+, Fe3+, Zn2+, V5+, CO32-, HCO3ˉ, Clˉ, SO42-, PO43- and humic acid (HA). The

concentrations of these ions were taken according to their WHO permissible limits in

drinking water [38].

The adsorption capacities (qe ) of uranyl ion in the presence of different interfering

ions are given in Fig. 6. The corresponding UV-visible spectra of the respective batches of

14
residual uranyl ions together with interfering cations and anions present in supernatant and

complexed with Arsenazo (III) are given as supporting information (Fig. S8a and 8b,

respectively). Except Ca2+ and Mg2+, the other cations and anions did not interfere with the

adsorption capacity of uranium by CARGO-1. The marginal decrease in the qe values for

Ca2+ and Mg2+ (i.e., 162.7 mg/g and 167 mg/g) as interfering cations, respectively, is

attributable to their competing nature with the UO22+ for the available binding sites at

CARGO-1. Similarly, the adsorption capacities (qe) of uranium decreased in the presence of

CO32- and HCO3- as interfering anions (154.5 mg/g and 155.9 mg/g, respectively). In this

case, CO32- and HCO3- could favor formation of neutral or negatively charged complexes of

uranium, which were electrostatically unfavorable for binding with negatively charged sites

in CARGO-1.

3.2.3 Adsorption kinetics

From the qt vs. t (contact time) plot (Fig. 7a), the uranium adsorption on CARGO-1

was non-linear. The adsorption was rapid during the first 10 minutes (i.e., qt = 148 mg/g in 10

min) and then equilibrium was reached in 90 min. The initial faster adsorption of uranium

ions is attributed to availability of higher concentrations of exposed binding sites on a flat

two dimensional surface of CARGO-1. Owing to faster initial rate of adsorption, the

adsorption data was fitted with linear forms of pseudo first order and pseudo second order

kinetic models. The linear form of pseudo first order kinetic model given as [55]:

1𝑘
log(𝑞𝑒 − 𝑞𝑡 ) = 𝑙𝑜𝑔𝑞𝑒 − 2.303 𝑡 (A3)

where k1 is the pseudo first order rate constant (in min-1), t is time (in min), qe and qt are

adsorption capacities (in mg/g) at equilibrium and at any time t, respectively. The linear

fitting of the plot of log(qe-qt) vs. t is poor (R2 = 0.93, Fig. 7b) and the theoretical adsorption

capacity calculated from the linear fit (qe = 51.28 mg/g) was not in agreement with the

experimental qe = 175.1 mg/g. So it can be commented that the rate of uranyl ion adsorption
15
by CARGO-1 was not simply dependent on the adsorption capacity of CARGO-1. The linear

form of the pseudo second order model is given as [56]:

𝑡 1 1
= − 𝑡 (A4)
𝑞𝑡 𝑘2 𝑞𝑒2 𝑞𝑒

Where k2 is the second order rate constant, t is time (in min) and qe and qt represent

adsorption capacities (in mg/g) at equilibrium and at any time t, respectively. The plot of t/qt

vs. 1/t revealed an excellent linear fit (R2 = 0.999, Fig. 7c), suggesting that the adsorption of

uranyl ions on CARGO-1 is a chemisorption process involving valency forces through

sharing or exchange of electrons between adsorbate and adsorbent [57]. From the slope of the

pseudo second order kinetics plot, the qe value was calculated as 176.36 mg/g, which is in

good agreement with the experimental qe = 175.13 mg/g. The pseudo second order rate

constant, k2 was calculated from the intercept of the plot and was found to be 2.92x10-3 g. mg-
1
. min-1. Such small k2 indicated decrease in the uptake of uranium species by the CARGO-1

with time due to decrease in the number of available binding sites [57].

Furthermore, the kinetic adsorption data was also tested with the Weber-Morris intra-

particle diffusion model, given as below [58]:

𝑞𝑡 = 𝐾𝑖 𝑡1/2 + 𝐶 (A5)

where qt is adsorption capacity (in mg/g) at any time t, Ki is intraparticle diffusion rate

constant (in mg.g−1.min−1/2), t is time (in min) and C is intercept. The intraparticle diffusion

plot between qt and t1/2 for the adsorption of uranyl ions on the surface of CARGO-1 is shown

in Fig. 7d. The plot revealed two linear profiles, which indicated that the adsorption of uranyl

ions on the surface of CARGO-1 occurred in two steps. The adsorption of uranyl ions on

CARGO-1 was not based on diffusion mechanism as the plot did not pass through the origin.

The intra-particle rate constant, ki is determined from the slope of the linearly fitted region

and was found to be 5.41 and 1.22 mg.g-1.min-1/2 for region I and II, respectively. The

decrease in ki values with time indicated slowing down of the uranyl ion adsorption. This is

16
attributable to coverage of active binding sites by uranyl ions with time. The parameter C for

the two regions was determined from the intercept of the linearly fitted portions, which were

128.76 and 161.66, respectively. The enhancement in the C value with time indicated

increase in the thickness of the boundary layer formed on the surface of CARGO-1 adsorbent

during the adsorption process.

3.2.4. Adsorption isotherms

The mechanism of uranyl ion adsorption on CARGO-1 was studied with respect to

Langmuir and Freundlich adsorption isotherm model in the temperature range of 288 K and

313 K. The initial concentration of the uranyl ions (Co) were taken in the range 10-250 mg/L.

At the optimized conditions, the equilibrium concentrations (Ce) of uranyl ions corresponding

to the respective initial concentrations (Co) of uranyl ions were measured and the respective

qe values were determined using the equation A2. The adsorption of uranyl ions on a

heterogeneous surface at the solid-liquid interface was studied using a non-linear Freundlich

adsorption isotherm model, which is expressed as [59]:

1/𝑛
𝑞𝑒 = 𝐾𝐹 𝐶𝑒 (A6)

Here, KF and n are the Freundlich isotherm constants and is related to adsorption capacity and

adsorption intensity, respectively. Fig. 8a represents non linear fit of the experimental data.

The fitting was done using the equation 𝑦 = 𝑎𝑥 𝑏 where the parameters a & b correspond to

KF and 1/n, respectively. The experimental data did not fit well with Freundlich adsorption

isotherm model as suggested from the poor R2 values provided in Table 1, and hence it may

be remarked that the uranyl ion adsorption on CARGO-1 was not governed by multilayer

adsorption phenomenon as proposed for Freundlich isotherm condition.

Next, the adsorption isotherm data was studied with respect to monolayer adsorption

proposed in Langmuir adsorption isotherm, given as [60,61]:

17
𝑞𝑚𝑎𝑥 𝐶𝑒
𝑞𝑒 = 1 (A7)
+𝐶𝑒
𝐾𝐿

The equilibrium parameter, RL is given as:

1
𝑅𝐿 = (A8)
1+(1+𝐾𝐿 𝐶𝑂 )

Here, qmax is the maximum theoretical adsorption capacity (mg/g), KL is the Langmuir

equilibrium constant (L/mg) and RL is the separation factor. Fig. 8b represents the plot of qe

𝑃 𝑥
vs. Ce, which was fitted by non-linear hyperbolic mathematical expression: 𝑦 = 𝑃 1+𝑥. By
2

comparing this expression with A7, P1 corresponds to qmax and P2 corresponds to 1/KL. The

R2, qmax and KL values obtained from the non-linear fit are given in Table 1. The R2 values

suggested that the adsorption of uranyl ions on CARGO-1 was primarily based on monolayer

adsorption phenomenon. As reflected from a comprehensive list of several types of reduced

graphene oxide adsorbents for uranyl ion adsorption (Table 2), the qmax for CARGO

adsorbent was determined to be 337.93 mg/g, which was significantly improved, particularly

among the graphene oxides reduced by simple ligands [21,24-31,48,54,56,57,62-88].

Furthermore, the qmax values increased with the temperature of the reaction medium, which

favored the adsorption phenomenon [87]. Similarly, the favorable adsorption process was

also marked from the RL values, which were between 0 and 1.

3.2.5. Adsorption thermodynamics and mechanism

The pseudo second order fitting of the adsorption kinetic data implied that the uranyl ion

adsorption on CARGO is governed by chemisorption phenomenon. The thermodynamics of

the adsorption process provides better insight about the mechanism of adsorption. In this

regard, the Gibb’s free energy at equilibrium condition (ΔGº) was calculated from the

Langmuir constant (KL) by the following equation [89-91]:

∆𝐺 𝑜 = −𝑅𝑇𝑙𝑛(𝐾𝐿 ∗ 𝑀. 𝑊. ) (A9)

18
Here the KL was converted form L/mg to L/mol by multiplying with the molecular weight of

uranyl ions. The values of thermodynamic constants, ΔHº and ΔSº were determined from the

slope and intercept of the plot of lnKL vs. 1/T, respectively (Fig. 9). The ΔHº value was

determined as 2.43 kJ/mol and ΔSº was found to be 92.53 J/mol. The ΔGº was negative and

increased with temperature, i.e., -24.2 kJ/mol at 288 K, -25.2 kJ/mol at 298 K and -26.5

kJ/mol at 313 K, which implied that the adsorption of uranyl ions on CARGO-1 was

thermodynamically favourable.

The adsorption of uranyl ions by CARGO-1 nanosheets is evident from the FE-SEM studies

(Fig. 10a). The EDX spectrum revealed M X-ray of uranium (Fig. 10b). The elemental maps

of the batch of CARGO-1 nanosheets after adsorbing uranyl ions revealed uranium

distribution (Fig. 10c) spatially correlated with that of C map (Fig. 10d) and N map (Fig. 10e.

The enhanced uranyl ion adsorption on CARGO-1 was attributable to several parameters: (a)

increase in the surface area; (b) suitable negative surface charge over a wide pH region; (c)

increase in the amide functionalities in CARGO. The first two parameters were

experimentally confirmed. The role of amide functionalities can be conclusively proved.

Firstly, there was marked increase in the qe value of uranyl ion adsorption on CARGO as

compared to GO. It implied that only carboxylic acid group is not sufficient for enhancing

uranyl ion adsorption. Further, ligands containing amines and amides have strong affinity for

binding with uranyl ions. In the present study, L-cysteine was modified to L-cysteine amide

to increase the amine and amide groups in each ligand. When GO was reduced by L-cysteine

amide then two nitrogen bearing functional groups were anchored to reduced GO to form

CARGO. Consequently, uranium adsorption on CARGO was favorable and followed

chemisorptions mechanism as suggested by adsorption isotherm studies and by pseudo

second order kinetic studies.

3.2.6. Regeneration and reusability

19
The reusability of CARGO-1 as an adsorbent for uranyl ions was also studied. In the

first cycle, 87.9 % adsorption of uranyl ions was achieved while 93.1 % of the adsorbed

uranyl ions were desorbed. The spent adsorbent was washed with 25 mL of 0.1 M HNO3

followed by conditioning with 25 mL 0.1 M liq. NH3 for the regeneration of the active biding

sites. The FT-IR spectrum of the re-generated adsorbent is given as supporting information

Fig. S9 which exhibited a doublet at 3426 cm-1 and 3114 cm-1 for N–H stretching vibrations.

It also exhibited high intensity peaks at 1624 cm-1 for the carbonyl group of amide and at

1401 cm-1 for C–H vibrations modes. In the subsequent cycles the percentage adsorption of

uranyl ions was 86.5, 85, 79.4 and 81.7 %, respectively. After five cycles, the adsorption

capacity decreased from 176 mg/g to 163 mg/g. Hence, it can be concluded that CARGO-1

can be used multiple times as an efficient adsorbent for removing high concentrations of

uranyl ions without sacrificing its adsorption capacity. There exists enough scope for

modifying the CARGO batch to facilitate removal of other uranium species that exists in

neutral pH so that the adsorbent can be used for uranium removal in drinking water.

4. Conclusion

A novel graphene based nano-adsorbent has been developed by reducing graphene

oxide (GO) with L-cysteine amide to introduce multiple amine/amide groups for enhanced

adsorption of uranium from aqueous medium. The L-cysteine amide as a precursor has been

synthesized and confirmed from FT-IR and 1H NMR. The reduction of GO by L-cysteine

amide (referred here as CARGO) has been characterized by XRD, UV-visible, FT-IR and

Raman spectroscopy. The batch of CARGO-1 prepared by treating 1:1 weight ratio of L-

cysteine amide and GO revealed an optimized 89 % uranyl ion adsorption from aqueous

medium (pH 5) in 60 min. The selectivity of uranyl ion adsorption was high, with a mild

interference from Ca2+, Mg2+, HCO3ˉ and CO32-. The kinetic and thermodynamic studies

favored adsorption of uranyl ions on CARGO where pseudo second order kinetic model and

20
Langmuir adsorption isotherm model were followed. The adsorption was attributed to

chemisorption phenomenon with qmax of 337.93 mg/g. Such high adsorption capacity was due

to affinity of uranyl ions with the amine/amide groups in CARGO adsorbent. A

comprehensive list of various graphene based adsorbents has been discussed, which revealed

that the qmax value obtained in the present study was among the higher ones. The enhanced

adsorption has been attributed to (a) increase in multiple amines/amides binding sites on the

graphene structure; (b) increase in BET surface area; and (c) negative surface charge

favorable for electrostatic interaction with uranyl ions.

Acknowledgement

Swati Verma wishes to thank Council for Scientific and Industrial Research (CSIR),

Government of India for providing Senior Research Fellowship. R.K. Dutta is thankful to

Board of Research in Nuclear Science (BRNS) Government of India, for sanctioning research

grant (2011/37C/37/BRNS) to pursue this work. Authors are thankful to Institute

Instrumentation Centre, IIT Roorkee for XRD and TEM facilities.

21
References:

[1] E.S. Craft, A.W. Abu-Qare, M.M. Flaherty, M.C. Garofolo, H.L. Rincavage, M.B.

Abou-Donia, Depleted and Natural Uranium: Chemistry and Toxicological Effects, J.

Toxicol. Environ. Heal. Part B. 7 (2004) 297–317. doi:10.1080/10937400490452714.

[2] B. Gu, Y.K. Ku, P.M. Jardine, Sorption and binary exchange of nitrate, sulfate, and

uranium on an anion-exchange resin, Environ. Sci. Technol. 38 (2004) 3184–3188.

doi:10.1021/es034902m.

[3] J. Shen, A. Schäfer, Removal of fluoride and uranium by nanofiltration and reverse

osmosis: A review, Chemosphere. 117 (2014) 679–691.

doi:10.1016/j.chemosphere.2014.09.090.

[4] T. Gafvert, C. Ellmark, E. Holm, Removal of radionuclides at a waterworks, J.

Environ. Radioact. 63 (2002) 105–115. doi:10.1016/S0265-931X(02)00020-6.

[5] C. Kütahyalı, M. Eral, Selective adsorption of uranium from aqueous solutions using

activated carbon prepared from charcoal by chemical activation, Sep. Purif. Technol.

40 (2004) 109–114. doi:10.1016/j.seppur.2004.01.011.

[6] R. Donat, The removal of uranium (VI) from aqueous solutions onto natural sepiolite,

J. Chem. Thermodyn. 41 (2009) 829–835. doi:10.1016/j.jct.2009.01.009.

[7] S. Aytas, M. Yurtlu, R. Donat, Adsorption characteristic of U(VI) ion onto thermally

activated bentonite, J. Hazard. Mater. 172 (2009) 667–674.

doi:10.1016/j.jhazmat.2009.07.049.

[8] A. Mellah, S. Chegrouche, M. Barkat, The removal of uranium (VI) from aqueous

solutions onto activated carbon: Kinetic and thermodynamic investigations, J. Colloid

Interface Sci. 296 (2006) 434–441. doi:10.1016/j.jcis.2005.09.045.

[9] R. Qadeer, J. Hanif, M. Saleem, M. Afzal, Effect of alkali metals, alkaline earth

metals and lanthanides on the adsorption of uranium on activated charcoal from

aqueous solutions, J. Radioanal. Nucl. Chem. Lett. 165 (1992) 243–253.

22
doi:10.1007/BF02164763.

[10] H. Sepehrian, M. Samadfam, Z. Asadi, Studies on the recovery of uranium from

nuclear industrial effluent using nanoporous silica adsorbent, Int. J. Environ. Sci.

Technol. 9 (2012) 629–636. doi:10.1007/s13762-012-0065-3.

[11] P. Misaelides, A. Godelitsas, A. Filippidis, D. Charistos, I. Anousis, Thorium and

uranium uptake by natural zeolitic materials, Sci. Total Environ. 173–174 (1995)

237–246. doi:10.1016/0048-9697(95)04748-4.

[12] I.I. Fasfous, J.N. Dawoud, Uranium (VI) sorption by multiwalled carbon nanotubes

from aqueous solution, Appl. Surf. Sci. 259 (2012) 433–440.

doi:10.1016/j.apsusc.2012.07.062.

[13] F.-L. Fan, Z. Qin, J. Bai, W.-D. Rong, F.-Y. Fan, W. Tian, X.-L. Wu, Y. Wang, L.

Zhao, Rapid removal of uranium from aqueous solutions using magnetic Fe3O4@SiO2

composite particles., J. Environ. Radioact. 106 (2012) 40–6.

doi:10.1016/j.jenvrad.2011.11.003.

[14] M. Wazne, X. Meng, G.P. Korfiatis, C. Christodoulatos, Carbonate effects on

hexavalent uranium removal from water by nanocrystalline titanium dioxide, J.

Hazard. Mater. 136 (2006) 47–52. doi:10.1016/j.jhazmat.2005.11.010.

[15] A. Nilchi, T.S. Dehaghan, S.R. Garmarodi, Kinetics , isotherm and thermodynamics

for uranium and thorium ions adsorption from aqueous solutions by crystalline tin

oxide nanoparticles, Desalination. 321 (2013) 67–71.

doi:10.1016/j.desal.2012.06.022.

[16] W. Li, J.T. Mayo, D.N. Benoit, L.D. Troyer, Z.A. Lewicka, B.J. Lafferty, J.G.

Catalano, S.S. Lee, V.L. Colvin, J.D. Fortner, Engineered superparamagnetic iron

oxide nanoparticles for ultra-enhanced uranium separation and sensing, J. Mater.

Chem. A. 4 (2016) 15022–15029. doi:10.1039/C6TA04709B.

[17] T.S. Anirudhan, J.R. Deepa, Binusreejayan, Synthesis and characterization of multi-

23
carboxyl-functionalized nanocellulose/nanobentonite composite for the adsorption of

uranium (VI) from aqueous solutions: Kinetic and equilibrium profiles, Chem. Eng. J.

273 (2015) 390–400. doi:10.1016/j.cej.2015.03.007.

[18] D. Humelnicu, C. Blegescu, D. Ganju, Removal of uranium (VI) and thorium (IV)

ions from aqueous solutions by functionalized silica: Kinetic and thermodynamic

studies, J. Radioanal. Nucl. Chem. 299 (2014) 1183–1190. doi:10.1007/s10967-013-

2873-4.

[19] T.S. Anirudhan, J. Nima, S. Sandeep, V.R.N. Ratheesh, Development of an amino

functionalized glycidylmethacrylate-grafted-titanium dioxide densified cellulose for

the adsorptive removal of arsenic (V) from aqueous solutions, Chem. Eng. J. 209

(2012) 362–371. doi:10.1016/j.cej.2012.07.129.

[20] G. Zhao, J. Li, X. Ren, C. Chen, X. Wang, Few-layered graphene oxide nanosheets

for heavy metal ion pollution management, Environ. Sci. Technol. 45 (2011) 10454–

10462. doi:dx.doi.org/10.1021/es203439v.

[21] Z. Li, F. Chen, L. Yuan, Y. Liu, Y. Zhao, Z. Chai, W. Shi, Uranium (VI) adsorption

on graphene oxide nanosheets from aqueous solutions, Chem. Eng. J. 210 (2012)

539–546. doi:10.1016/j.cej.2012.09.030.

[22] R. Sitko, E. Turek, B. Zawisza, E. Malicka, E. Talik, J. Heimann, A. Gagor, B. Feist,

R. Wrzalik, Adsorption of divalent metal ions from aqueous solutions using graphene

oxide, Dalt. Trans. 42 (2013) 5682–5689. doi:10.1039/c3dt33097d.

[23] G. Benay, G. Wipff, Liquid-Liquid Extraction of Uranyl by an Amide Ligand:

Interfacial Features Studied by MD and PMF Simulations, J. Phys. Chem. B. 117

(2013) 7399–7415. doi:10.1021/jp4028386.

[24] F. Wang, H. Li, Q. Liu, Z. Li, R. Li, H. Zhang, L. Liu, G.A. Emelchenko, J. Wang, A

graphene oxide/amidoxime hydrogel for enhanced uranium capture, Sci. Rep. 6

(2016) 19367. doi:10.1038/srep19367.

24
[25] Y. Wang, Z. Wang, R. Ang, J. Yang, N. Liu, J. Liao, Y. Yang, J. Tang, Synthesis of

amidoximated graphene oxide nanoribbons from unzipping of multiwalled carbon

nanotubes for selective separation of uranium (VI), RSC Adv. 5 (2015) 89309–89318.

doi:10.1039/c5ra15977f.

[26] Y. Zhao, J. Li, S. Zhang, H. Chen, D. Shao, Efficient enrichment of uranium (VI) on

amidoximated magnetite/graphene oxide composites, RSC Adv. 3 (2013) 18952–

18959. doi:10.1039/c3ra42236d.

[27] S. Liu, J. Ma, W. Zhang, F. Luo, M. Luo, F. Li, L. Wu, Three-dimensional graphene

oxide/phytic acid composite for uranium (VI) sorption, J. Radioanal. Nucl. Chem. 306

(2015) 507–514. doi:10.1007/s10967-015-4162-x.

[28] R. Hu, D. Shao, X. Wang, Graphene oxide/polypyrrole composites for highly

selective enrichment of U (VI) from aqueous solutions, Polym. Chem. 5 (2014) 6207–

6215. doi:10.1039/C4PY00743C.

[29] Z. Zhao, J. Li, T. Wen, C. Shen, X. Wang, A. Xu, Surface functionalization graphene

oxide by polydopamine for high affinity of radionuclides, Colloids Surfaces A

Physicochem. Eng. Asp. 482 (2015) 258–266. doi:10.1016/j.colsurfa.2015.05.020.

[30] Y. Sun, D. Shao, C. Chen, S. Yang, X. Wang, Highly E fficient Enrichment of

Radionuclides on Graphene Oxide - Supported Polyaniline, Environ. Sci. Technol. 47

(2013) 9904–9910. doi:10.1017/CBO9781107415324.004.

[31] W. Cheng, M. Wang, Z. Yang, Y. Sun, C. Ding, The efficient enrichment of U (VI)

by graphene oxide-supported chitosan, RSC Adv. 4 (2014) 61919–61926.

doi:10.1039/C4RA09541C.

[32] D.C. Marcano, D. V. Kosynkin, J.M. Berlin, A. Sinitskii, Z. Sun, A. Slesarev, L.B.

Alemany, W. Lu, J.M. Tour, Improved synthesis of graphene oxide, ACS Nano. 4

(2010) 4806–4814. doi:10.1021/nn1006368.

[33] H.P. Boehm, Some aspects of the surface chemistry of carbon blacks and other

25
carbons, Carbon. 32 (1994) 759–769. doi:10.1016/0008-6223(94)90031-0.

[34] T.A. Saleh, The influence of treatment temperature on the acidity of MWCNT

oxidized by HNO3 or a mixture of HNO3/H2SO4, Appl. Surf. Sci. 257 (2011) 7746–

7751. doi:10.1016/j.apsusc.2011.04.020.

[35] I. Feliciano-Ramos, M. Caban-Acevedo, M. Aulice Scibioh, C.R. Cabrera, Self-

assembled monolayers of L-cysteine on palladium electrodes, J. Electroanal. Chem.

650 (2010) 98–104. doi:10.1016/j.jelechem.2010.09.001.

[36] Z. Bujňáková, M. Baláž, E. Dutková, P. Baláž, M. Kello, G. Mojžišová, J. Mojžiš, M.

Vilková, J. Imrich, M. Psotka, Mechanochemical approach for the capping of mixed

core CdS/ZnS nanocrystals: Elimination of cadmium toxicity, J. Colloid Interface Sci.

486 (2017) 97–111. doi:10.1016/j.jcis.2016.09.033.

[37] G. Song, D. Chen, C.L. Pan, R.H. Crabtree, X. Li, Rh-catalyzed oxidative coupling

between primary and secondary benzamides and alkynes: Synthesis of polycyclic

amides, J. Org. Chem. 75 (2010) 7487–7490. doi:10.1021/jo101596d.

[38] 4th Ed., WHO guidelines for drinking-water quality., 2003. doi:10.1016/S1462-

0758(00)00006-6.

[39] L. Jauberty, N. Drogat, J.-L. Decossas, V. Delpech, V. Gloaguen, V. Sol,

Optimization of the arsenazo-III method for the determination of uranium in water

and plant samples., Talanta. 115 (2013) 751–754. doi:10.1016/j.talanta.2013.06.046.

[40] S. Gilje, R.B. Kaner, G.G. Wallace, D.A.N. Li, M.B. Mu, M.B. Muller, S. Gilje, R.B.

Kaner, G.G. Wallace, Processable aqueous dispersions of graphene nanosheets, Nat.

Nanotechnol. 3 (2008) 101–105. doi:10.1038/nnano.2007.451.

[41] D. Chen, L. Li, L. Guo, An environment-friendly preparation of reduced graphene

oxide nanosheets via amino acid., Nanotechnology. 22 (2011) 325601.

doi:10.1088/0957-4484/22/32/325601.

[42] J. Zhang, H. Yang, G. Shen, P. Cheng, J. Zhang, S. Guo, Reduction of graphene oxide

26
via L-ascorbic acid., Chem. Commun. 46 (2010) 1112–1114. doi:10.1039/b917705a.

[43] K. Krishnamoorthy, M. Veerapandian, K. Yun, S.J. Kim, The chemical and structural

analysis of graphene oxide with different degrees of oxidation, Carbon. 53 (2013) 38–

49. doi:10.1016/j.carbon.2012.10.013.

[44] S. Park, J. An, I. Jung, R.D. Piner, S.J. An, X. Li, A. Velamakanni, R.S. Ruoff,

Colloidal suspensions of highly reduced graphene oxide in a wide variety of organic

solvents., Nano Lett. 9 (2009) 1593–1597. doi:10.1021/nl803798y.

[45] S. Park, J. An, J.R. Potts, A. Velamakanni, S. Murali, R.S. Ruoff, Hydrazine-

reduction of graphite- and graphene oxide, Carbon. 49 (2011) 3019–3023.

doi:10.1016/j.carbon.2011.02.071.

[46] H.J. Shin, K.K. Kim, A. Benayad, S.M. Yoon, H.K. Park, I.S. Jung, M.H. Jin, H.K.

Jeong, J.M. Kim, J.Y. Choi, Y.H. Lee, Efficient reduction of graphite oxide by

sodium borohydride and its effect on electrical conductance, Adv. Funct. Mater. 19

(2009) 1987–1992. doi:10.1002/adfm.200900167.

[47] J.I. Parades, S. Villar-Rodil, A. Martínez-Alonso, J.M.D. Tascón, Graphene oxide

dispersions in organic solvents, Langmuir. 24 (2008) 10560–10564.

doi:10.1021/la801744a.

[48] S. Verma, R.K. Dutta, A facile method of synthesizing ammonia modified graphene

oxide for efficient removal of uranyl ions from aqueous medium, RSC Adv. 5 (2015)

77192–77203. doi:10.1039/C5RA10555B.

[49] T.A. Saleh, A. Sari, M. Tuzen, Effective adsorption of antimony(III) from aqueous

solutions by polyamide-graphene composite as a novel adsorbent, Chem. Eng. J. 307

(2017) 230–238. doi:10.1016/j.cej.2016.08.070.

[50] E. Toral-Sánchez, J.A. Ascacio Valdés, C.N. Aguilar, F.J. Cervantes, J.R. Rangel-

Mendez, Role of the intrinsic properties of partially reduced graphene oxides on the

chemical transformation of iopromide, Carbon. 99 (2016) 456–465.

27
doi:10.1016/j.carbon.2015.12.067.

[51] T.A. Saleh, M.M. Al-Shalalfeh, A.A. Al-Saadi, Graphene Dendrimer-stabilized silver

nanoparticles for detection of methimazole using Surface-enhanced Raman scattering

with computational assignment, Sci. Rep. 6 (2016) 32185. doi:10.1038/srep32185.

[52] K.N. Kudin, B. Ozbas, H.C. Schniepp, R.K. Prud’homme, I.A. Aksay, R. Car, Raman

Spectra of Graphite Oxide and Functionalized Graphene Sheets, Nano Lett. 8 (2007)

36–41. doi:10.1021/nl071822y.

[53] C. Moulin, P. Decambox, V. Moulin, J.G. Decaillon, Uranium Speciation in Solution

by Time-Resolved Laser-Induced Fluorescence, Anal. Chem. 67 (1995) 348–353.

doi:10.1021/ac00098a019.

[54] G. Zhao, T. Wen, X. Yang, S. Yang, J. Liao, J. Hu, D. Shao, X. Wang,

Preconcentration of U (VI) ions on few-layered graphene oxide nanosheets from

aqueous solutions, Dalt. Trans. 41 (2012) 6182–6188. doi:10.1039/c2dt00054g.

[55] Y.S. Ho, G. McKay, Pseudo-second order model for sorption processes, Process

Biochem. 34 (1999) 451–465. doi:10.1016/S0032-9592(98)00112-5.

[56] P. Zong, S. Wang, Y. Zhao, H. Wang, H. Pan, C. He, Synthesis and application of

magnetic graphene/iron oxides composite for the removal of U(VI) from aqueous

solutions, Chem. Eng. J. 220 (2013) 45–52. doi:10.1016/j.cej.2013.01.038.

[57] L. Chen, D. Zhao, S. Chen, X. Wang, C. Chen, One-step fabrication of amino

functionalized magnetic graphene oxide composite for uranium (VI) removal., J.

Colloid Interface Sci. 472 (2016) 99–107. doi:10.1016/j.jcis.2016.03.044.

[58] J.C. Weber, W.J. Morris, Kinetics of Adsorption on Carbon from Solution, J. Sanit.

Eng. Div. 89 (1963) 31–60.

[59] H.M.F. Freundlich, Over the adsorption in solutions, J. Phys. Chem. 57 (1906) 385–

471.

[60] E.C. Lima, M.A. Adebayo, F.M. Machado, Chapter 3: Kinetic and Equilibrium

28
Models of Adsorption. Springer, 2015.

[61] I. Langmuir, The Adsorption of Gases on Plane Surfaces of Glass, Mica and

Platinum, J. Am. Chem. Soc. 40 (1918) 1361–1403. doi:10.1021/ja02242a004.

[62] H. Meng, Z. Li, F. Ma, X. Wang, W. Zhou, L. Zhang, Synthesis and characterization

of surface ion-imprinted polymer based on SiO2-coated graphene oxide for selective

adsorption of uranium (VI), RSC Adv. 5 (2015) 67662–67668.

doi:10.1039/C5RA11149H.

[63] Y. Zhao, C. Guo, H. Fang, J. Jiang, Competitive adsorption of Sr (II) and U (VI) on

graphene oxide investigated by batch and modeling techniques, J. Mol. Liq. 222

(2016) 263–267. doi:10.1016/j.molliq.2016.07.032.

[64] C.L. Wang, Y. Li, C.L. Liu, Sorption of uranium from aqueous solutions with

graphene oxide, J. Radioanal. Nucl. Chem. 304 (2015) 1017–1025.

doi:10.1007/s10967-014-3855-x.

[65] Y. Sun, X. Wang, Y. Ai, Z. Yu, W. Huang, C. Chen, T. Hayat, A. Alsaedi, X. Wang,

Interaction of sulfonated graphene oxide with U (VI) studied by spectroscopic

analysis and theoretical calculations, Chem. Eng. J. (2016).

doi:10.1016/j.cej.2016.10.122.

[66] T. Hu, S. Ding, H. Deng, Application of three surface complexation models on U(VI)

adsorption onto graphene oxide, Chem. Eng. J. 289 (2016) 270–276.

doi:10.1016/j.cej.2015.12.030.

[67] X. Wang, Z. Chen, X. Wang, Graphene oxides for simultaneous highly efficient

removal of trace level radionuclides from aqueous solutions, Sci. China Chem. 58

(2015) 1766–1773. doi:10.1007/s11426-015-5435-5.

[68] L. Tan, J. Wang, Q. Liu, Y. Sun, X. Jing, L. Liu, J. Liu, D. Song, The synthesis of a

manganese dioxide – iron oxide – graphene magnetic nanocomposite for enhanced

uranium (VI) removal, New J. Chem. 39 (2015) 868–876. doi:10.1039/C4NJ01256A.

29
[69] W. Song, D. Shao, S. Lu, X. Wang, Simultaneous removal of uranium and humic acid

by cyclodextrin modified graphene oxide nanosheets, Sci. China Chem. 57 (2014)

1291–1299. doi:10.1007/s11426-014-5119-6.

[70] L. Shao, X. Wang, Y. Ren, S. Wang, J. Zhong, M. Chu, H. Tang, L. Luo, D. Xie,

Facile fabrication of magnetic cucurbit[6]uril/graphene oxide composite and

application for uranium removal, Chem. Eng. J. 286 (2016) 311–319.

doi:10.1016/j.cej.2015.10.062.

[71] Y. Sun, S. Yang, Y. Chen, C. Ding, W. Cheng, X. Wang, Adsorption and Desorption

of U (VI) on Functionalized Graphene Oxides: A Combined Experimental and

Theoretical Study, Environ. Sci. Technol. 49 (2015) 4255–4262.

doi:10.1021/es505590j.

[72] H. Cheng, K. Zeng, J. Yu, Adsorption of uranium from aqueous solution by graphene

oxide nanosheets supported on sepiolite, J. Radioanal. Nucl. Chem. 298 (2013) 599–

603. doi:10.1007/s10967-012-2406-6.

[73] N. Pan, L. Li, J. Ding, S. Li, R. Wang, Y. Jin, X. Wang, C. Xia, Preparation of

graphene oxide-manganese dioxide for highly efficient adsorption and separation of

Th(IV)/U(VI), J. Hazard. Mater. 309 (2016) 107–115.

doi:10.1016/j.jhazmat.2016.02.012.

[74] L.P. Lingamdinne, Y.L. Choi, I.S. Kim, J.K. Yang, J.R. Koduru, Y.Y. Chang,

Preparation and characterization of porous reduced graphene oxide based inverse

spinel nickel ferrite nanocomposite for adsorption removal of radionuclides, J.

Hazard. Mater. 326 (2017) 145–156. doi:10.1016/j.jhazmat.2016.12.035.

[75] C. Ding, W. Cheng, Y. Sun, X. Wang, Determination of chemical affinity of graphene

oxide nanosheets with radionuclides investigated by macroscopic, spectroscopic and

modeling techniques, Dalt. Trans. 43 (2014) 3888–3896. doi:10.1039/c3dt52881b.

[76] L. Tan, Q. Liu, D. Song, X. Jing, J. Liu, R. Li, S. Hu, L. Liu, J. Wang, Uranium

30
extraction using a magnetic CoFe2O4–graphene nanocomposite: kinetics and

thermodynamics studies, New J. Chem. 39 (2015) 2832–2838.

doi:10.1039/C4NJ01981D.

[77] S. Liu, S. Li, H. Zhang, L. Wu, L. Sun, J. Ma, Removal of uranium (VI) from aqueous

solution using graphene oxide and its amine-functionalized composite, J. Radioanal.

Nucl. Chem. 309 (2016) 607–614. doi:10.1007/s10967-015-4654-8.

[78] X. Liu, J. Li, X. Wang, C. Chen, X. Wang, High performance of phosphate-

functionalized graphene oxide for the selective adsorption of U(VI) from acidic

solution, J. Nucl. Mater. 466 (2015) 56–64. doi:10.1016/j.jnucmat.2015.07.027.

[79] Z. Wang, Y. Wang, J. Liao, Y. Yang, N. Liu, J. Tang, Improving the adsorption

ability of graphene sheets to uranium through chemical oxidation, electrolysis and

ball-milling, J. Radioanal. Nucl. Chem. 308 (2016) 1095–1102. doi:10.1007/s10967-

015-4598-z.

[80] L. Tan, Y. Wang, Q. Liu, J. Wang, X. Jing, L. Liu, J. Liu, D. Song, Enhanced

adsorption of uranium (VI) using a three-dimensional layered double

hydroxide/graphene hybrid material, Chem. Eng. J. 259 (2015) 752–760.

doi:10.1016/j.cej.2014.08.015.

[81] S. Chen, J. Hong, H. Yang, J. Yang, Adsorption of uranium (VI) from aqueous

solution using a novel graphene oxide-activated carbon felt composite, J. Environ.

Radioact. 126 (2013) 253–258. doi:10.1016/j.jenvrad.2013.09.002.

[82] L. Shao, J. Zhong, Y. Ren, H. Tang, X. Wang, Perhydroxy-CB[6] decorated graphene

oxide composite for uranium (VI) removal, J. Radioanal. Nucl. Chem. 311 (2017)

627–635. doi:10.1007/s10967-016-5067-z.

[83] Z.-B. Zhang, Y.-F Qiu, Y. Dai, P.-F Wang, B. Gao, Z.-M Dong, X.-H Cao, Y.-H Liu,

Z.-G Le, Synthesis and application of sulfonated graphene oxide for the adsorption of

uranium (VI) from aqueous solutions, J. Radioanal. Nucl. Chem. 310 (2016) 1–11.

31
doi:10.1007/s10967-016-4813-6.

[84] X. Song, L. Tan, X. Sun, H. Ma, L. Zhu, X. Yi, Q. Dong, J. Gao, Facile preparation of

NiCo2O4@rGO composites for the removal of uranium ions from aqueous solutions,

Dalt. Trans. 45 (2016) 16931–16937. doi:10.1039/C6DT03261C.

[85] H.H. El-Maghrabi, S.M. Abdelmaged, A.A. Nada, F. Zahran, S.A. El-Wahab, D.

Yahea, G.M. Hussein, M.S. Atrees, Magnetic graphene based nanocomposite for

uranium scavenging, J. Hazard. Mater. 322 (2017) 370–379.

doi:10.1016/j.jhazmat.2016.10.007.

[86] X. Wang, Q. Fan, S. Yu, Z. Chen, Y. Ai, Y. Sun, A. Hobiny, A. Alsaedi, X. Wang,

High sorption of U(VI) on graphene oxides studied by batch experimental and

theoretical calculations, Chem. Eng. J. 287 (2016) 448–455.

doi:10.1016/j.cej.2015.11.066.

[87] D. Shao, G. Hou, J. Li, T. Wen, X. Ren, X. Wang, PANI/GO as a super adsorbent for

the selective adsorption of uranium (VI), Chem. Eng. J. 255 (2014) 604–612.

doi:10.1016/j.cej.2014.06.063.

[88] Z.J. Li, L. Wang, L.Y. Yuan, C.L. Xiao, L. Mei, L.R. Zheng, J. Zhang, J.H. Yang,

Y.L. Zhao, Z.T. Zhu, Z.F. Chai, W.Q. Shi, Efficient removal of uranium from

aqueous solution by zero-valent iron nanoparticle and its graphene composite, J.

Hazard. Mater. 290 (2015) 26–33. doi:10.1016/j.jhazmat.2015.02.028.

[89] G.M. Barrow, Physical Chemistry (SiE), 5th ed., Tata McGraw-Hill Publishing

Company Limited, 2008.

[90] P.S. Ghosal, A.K. Gupta, Determination of thermodynamic parameters from

Langmuir isotherm constant-revisited, J. Mol. Liq. 225 (2017) 137–146.

doi:10.1016/j.molliq.2016.11.058.

[91] X. Zhou, X. Zhou, The Unit Problem in the Thermodynamic Calculation of

Adsorption Using the Langmuir Equation, Chem. Eng. Commun. 201 (2014) 1459–

32
1467. doi:10.1080/00986445.2013.818541.

FIGURE CAPTION

Fig. 1 FT-IR spectra of (a) GO; (b) CARGO-1; (c) CARGO-2 and (d) CARGO-3.

Fig. 2 Raman spectra of (a) GO; (b) CARGO-1; (c) CARGO-2 and (d) CARGO-3.

Fig. 3 Transmission electron microscopy image of CARGO-1 showing flake like nanosheet

structure

Fig. 4 Showing comparison of adsorption kinetics of uranyl ions on GO and CARGO for Co =

100 mg/L, m = 50 mg, V = 100 mL.

Fig. 5 Showing (a) efficiency of adsorption of uranyl ions by CARGO-1, CARGO-2 and

CARGO-3 adsorbents for Co = 100 mg/L, m = 10-100 mg, V = 100 mL; (b) effect of varying

doses of CARGO-1 on adsorption capacity (qe) and efficiency of uranyl ion adsorption for Co

= 100 mg/L, m = 10-100 mg, V = 100 mL; (c) effect of pH ranging between 2 and 9 on the qe

values of uranyl ions adsorption on CARGO-1 for Co = 100 mg/L, m = 50 mg, V = 100 mL

and (d) zeta potential measurements of CARGO-1 in the pH range 2-9. The error bars in Fig.

5a and 5b correspond to standard deviation of triplicate measurements.

Fig. 6 Showing the effect of interfering ions on the adsorption capacity (qe) for Co = 100

mg/L, m = 50 mg, V = 100 mL.

Fig. 7 Showing (a) the effect of contact time on adsorption capacity (qe) of uranyl ions

adsorption on CARGO-1 for Co = 100 mg/L, m = 50 mg, V = 100 mL, t = 2-120 min; (b)

adsorption of uranyl ions on CARGO-1 fitted with pseudo first order kinetic model; (c)

adsorption of uranyl ions on CARGO-1 fitted with pseudo second order model and (d) intra-

particle diffusion plot for the adsorption of uranyl ions.

33
Fig. 8 Adsorption of uranyl ions on CARGO-1 modeled by: (a) non-linear Freundlich

adsorption isotherm and (b) Langmuir adsorption isotherm at pH 5 and at different

temperatures, e.g., 288 K, 298 K & 313 K. The experimental conditions were Co = 10-250

mg/L, m = 50 mg, V = 100 mL.

Fig. 9 Van’t Hoff plot between lnKL and 1/T.

Fig. 10 Representative scanning electron microscopy image of (a) CARGO-1 after uranyl

ions adsorption; (b) corresponding EDX spectrum; (c) elemental map of uranium; (d)

elemental map of carbon; (e) elemental map of oxygen and (f) elemental map of nitrogen.

34
FIGURES

3438 3023 1584


N-H C-H 1623 N-H
2584
C=O 1486 1404
(d) 3140 S-H C=C C-H
Transmittance O-H

(c)

(b)

(a)

1736
1637 1222 1050
3420 C=O
C=C C-O C-O-C
O-H
4000 3400 2800 2200 1600 1000 400
-1
Wavenumber (cm )

Fig. 1

35
(d) 1352 1599

ID/IG = 0.917

(c) 1355 1601

ID/IG = 0.918
Intensity

(b) 1357 1604

ID/IG = 0.901

(a) 1363 1604

ID/IG = 0.822

1000 1200 1400 1600 1800 2000


-1
Raman shift (cm )

Fig. 2

36
Fig. 3

37
200

160

qt (mg/g) 120

80

40
CARGO-1
GO
0
0 20 40 60 80 100 120
Time (min)

Fig. 4

38
100
(a)
80

% Removal 60

40

20 CARGO-1
CARGO-2
CARGO-3
0
0 20 40 60 80 100
Adsorbent dose (mg)

600
(b) CARGO-1
500 CARGO-2
CARGO-3
400
qe (mg/g)

300

200

100

0
0 20 40 60 80 100
Adsorbent dose (mg)

Fig. 5 (continue to the next page)

39
190
(c)
170

qe (mg/g)
150

130

110

90
1 2 3 4 5 6 7 8 9 10
pH

-27
(d)
-25
Zeta potential (mV)

-23

-21

-19

-17

-15
1 2 3 4 5 6 7 8 9 10
pH

Fig. 5

40
Fig. 6
qe (mg/g)

0
40
80
120
160
200

Con
trol
Na +
Ca 2 +
Mg 2 +

K+
Fe 2+
Fe 3+
Pb 2+
Zn 2+
V 5+
CO 2-
HC 3 -
O
3
SO 2-
4
Cl -
PO 3-
4
HA

41
180
(a)
170

qt (mg/g)
160

150

140

130
0 20 40 60 80 100 120
Time (min)

2.0
y = -0.019x + 1.71
2
1.5 R = 0.930

1.0
log (qe-qt)

0.5

0.0

-0.5

-1.0 (b)
0 20 40 60 80 100 120
Time (min)

Fig. 7 (continue to the next page)

42
0.8

0.7
(c)
0.6

0.5
t/qt

0.4

0.3

0.2

0.1 y = 0.0056x + 0.01


2
R = 0.999
0.0
0 20 40 60 80 100 120
Time (min)

180
(d)
170 Region II
qt (mg/g)

160
I
on

150
gi
Re

140

130
0 2 4 6 8 10 12
t1/2

Fig. 7

43
350
(a)
280

qe (mg/g)
210

140

70 288 K
298 K
313 K
0
0 20 40 60 80 100 120
Ce (mg/L)

350
(b)
280
qe (mg/g)

210

140

70 288 K
298 K
313 K
0
0 20 40 60 80 100 120
Ce (mg/L)

Fig. 8

44
10.22

10.20

10.18
ln KL
10.16

10.14

10.12

10.10
0.0032 0.0033 0.0034 0.0035
-1
1/T (K )

Fig. 9

45
Fig.10

46
NH2 NH2 NH2
SOCl2 liq. NH3 O
O O
50 OC,
SH OH 30 min SH Cl SH NH2

L-Cysteine acyl chloride Cysteine amide

OH OH HSH2C O OH OH O CONH2
H
HOOC COOH CH NH C C N CH
H2NOC CH2SH
Cysteine
O amide O
HOOC OH HO OH
microwave
treatment
O HSH2C O CONH2
H
HOOC COOH CH NH C C N CH
H2NOC O OH OH O CH2SH
OH OH

Graphene oxide CARGO

Scheme 1. Schematic representation and a proposed structure of the as synthesized cysteine

amide reduced graphene oxide.

47
Table 1 Freundlich and Langmuir adsorption isotherm constants for the adsorption of uranyl
ions by CARGO-1 nanosheets.

Freundlich constants Langmuir constants


Temperature qmax
KF (mg/g) 1/n R2 KL (L/mg) R2
(mg/g)
288 K 64.061 0.349 0.902 329.52 0.091 0.998
298 K 70.103 0.337 0.915 337.93 0.097 0.999
313 K 75.265 0.334 0.919 352.96 0.099 0.996

48
Table 2 List of graphene based adsorbents for uranium adsorption.
Contact
S. No. Adsorbent qmax (mg/g) pH Ref.
time (h)
1. GO/SiO2 17.89 4 0.66 62
2. GO 29.03 3 24 63
3. GO 40.22 3.5 24 64
4. Sulfonated GO 45.05 2 48 65
5. Fe3O4/GO 69.49 5 24 56
6. GO 76.92 4 24 66
7. GO 79.73 5.2 48 67
8. NH3-GO 80.13 6 4 48
9. MnO2-Fe3O4-rGO 95.24 6 6.667 68
10. Cyclodextrin modified GO 97.328 5 24 69
11. Few layered GO 97.5 5 24 54
12. Cucurbit[6]uril/GO-Fe3O4 122.48 5 24 70
13. GO/phytic acid 124.3 5.5 4 27
14. GOs 138.89 4 48 71
15. Amino functionalized magnetic GO 141.2 6 12 57
16. Polydopamine/GO 145.9 4 24 29
17. GO/ polypyrrole 147.1 5 48 28
18. GO@sepiolite 161.29 5 24 72
19. MnO2/GO 185.2 3.8 0.5 73
20. RGONF 200 3.5 3 74
21. GO 208.33 4 48 75
22. GO-chitosan 225.78 4 24 31
23. CoFe2O4-rGO 227.2 6 4 76
24. GO-NH2 227.3 5.5 6 77
25. Polyaniline@GO 245.14 3 48 30
26. Phosphate-GO 251.7 4 24 78
27. GO by chemical oxidation 257.23 4.5 2.5 79
28. 3-D layered double 277.80 4 3 80
hydroxide/graphene hybrid
29. Amidoximated magnetic GO 284.9 5 24 26
30. GO activated carbon felt 298 5.5 1 81
31. GO 299 4 24 21
32. Perhydroxy-CB[6]/GO 301.6 5 24 82
33. Sulfonated GO 309.09 6 3 83
34. NiCo2O4@rGO 333.3 5 9 84
35. GO/amidoxime hydro gel 398.41 6 7 24
36. Amidoximated GO 502.6 4.5 - 25
37. Ferberite-graphene 20 625 6 2 85
38. GO 1330 4.5 48 86
39. Polyaniline/GO 1960 5 48 87
40. Fe/rGO 4174 5 24 88
41. CARGO-1 337.93 5 2 This
study

49

Potrebbero piacerti anche