Sustainable Synthesis of MOF-5@GO Nanocomposites For E Cient Removal of Rhodamine B From Water

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

http://pubs.acs.

org/journal/acsodf Article

Sustainable Synthesis of MOF-5@GO Nanocomposites for Efficient


Removal of Rhodamine B from Water
Gyanendra Kumar and Dhanraj T. Masram*
Cite This: ACS Omega 2021, 6, 9587−9599 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: Herein, a metal−organic framework (MOF-5) is


synthesized by a solvothermal process and graphene oxide (GO) is
prepared from the improved Hummer’s method. The synthesis of
MOF-5@GO nanocomposites is one-pot process via a grinding
method and employed for the removal of Rhodamine B (RhB)
Downloaded via 103.120.38.88 on May 14, 2021 at 06:15:45 (UTC).

dye. The removal efficiency of RhB is found to be 60.64% (151.62


mg·g−1) at 500 ppm. About 98.88% of RhB is removed within 5
min of contact time and increased up to 99.68% up to 10 min. The
removal rate of MOF-5@GO nanocomposites is much better than
that of pristine MOF-5. Equilibrium adsorption capacity is
determined by a series of different experimental conditions such
as pH, time, and concentration of dye solution. Although the
results also showed that dye removal on MOF-5@GO nano-
composites is well described by the Langmuir isotherm (R2 = 0.9703), the adsorption kinetics data reveals pseudo-second-order (R2
= 0.9908). The synthesized nanocomposite is efficient for removal of dye, cost-effective, and reusable. Additionally, stability and self-
degradation studies of pure RhB are reported in aqueous solution for up to 120 days at different pH values (pH 1−12).

1. INTRODUCTION zole (HBT),10 alumina oxides,11 Fe3O4@L-arginine,12 porous


Rhodamine B (RhB) is a well-known cationic fluorescent dye silica beads,13 carbon nanotubes14 (CNT), carbon-based
that is widely used as a colorant in textiles, foodstuffs, scientific materials,15 and graphene oxide16 (GO); however, in these
research, and pharmaceutical industries.1 It is toxic and cases, there is poor sorption capacities, multi-step synthesis,
harmful to human beings and animals, causes damage to and low sorption rates. Therefore, MOFs/GO-based nano-
skin, eyes, and respiratory tract, and causes gene mutations. composites are potential material for the adsorption of dye due
Additionally, it is a well-known carcinogenic and neurotoxic to easy modification of surface, low toxicity, and large surface
material toward animals and humans.2 Worldwide, more than area. Consequently, there is an urgent need for these types of
10,000 textile dyes are used commercially with their yearly nanocomposite materials that can adsorb toxic dye more
production exceeding ∼7 × 105 metric tons,3 out of which efficiently. The cationic dye (RhB) is known to be better
about 2% (∼14 × 103 metric tons) of the annual dye adsorbed on the graphene oxide via strong interactions from
production is discharged as an effluent from manufacturing negatively charged oxygenated functional groups.17,18 Already
units. Color is one of the clearest indicators of water pollution available adsorbents are not only expensive but also difficult to
caused by the discharge of dye as a waste, which is responsible handle and redevelop after use. Meanwhile, metal−organic
for the damage to water bodies, aesthetic pollution, frameworks (MOFs) are highly ordered crystalline coordina-
eutrophication, and perturbations in aquatic life.4,5 RhB exists tion polymers and one of the most fascinating classes of
for a long time in a natural water body or aqueous medium. So, adsorbent,19 having the advantages of high surface area,
it is necessary to remediate dye from the aqueous solution. The porosity, excellent absorbability, good catalytic activity, and
remediation of dye from the aqueous medium is one of the controlled modification synthesis.20 Additionally, MOFs are
most challenging tasks. Among the several purification promising candidates for a variety of applications such as dye
techniques, adsorption process has been verified to be one of
the best and simplest techniques for eliminating pollutants
from water bodies. The advantages of the adsorption process Received: January 9, 2021
are its cost-effectiveness, accuracy, remarkable versatility, Accepted: March 17, 2021
simplicity in design, safety of operation, high efficiency, and Published: March 30, 2021
adsorbent reusability.6−8 There are many reports where a
variety of substrates are used for adsorption of dye from
wastewater such as Cu-MOF,9 modified 1-hydroxybenzotria-
© 2021 The Authors. Published by
American Chemical Society https://doi.org/10.1021/acsomega.1c00143
9587 ACS Omega 2021, 6, 9587−9599
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 1. FTIR spectra of (a) (i) GO, (ii) MOF-5, (iii) MOF-5@GO, (iv) RhB, and (v) RhB-MOF-5@GO nanocomposites. XRD pattern of (b)
MOF-5 standard and (c) (i) graphite powder, (ii) GO, (iii) MOF-5, (iv) MOF-5@GO, and (v) RhB-MOF-5@GO nanocomposites. (d, e) FTIR
and XRD of as-synthesized nanocomposites and after water immersion for up to 7 days.

adsorption, energy storage, catalysis, and gas separation.21 In problem due to lack of stability in water. The development of
the case of MOF-5, formerly described in 1999, it is made up MOF-GO composites occurs through interactions between
of ZnO4 units connected by a 1,4-benzenedicarboxylate linker oxygen groups of GO and the metallic centers of MOFs by
to form an amazing cubic network. The first series of MOF-5 chemical bonding or electrostatic interaction.40,41 Therefore, to
reported in 2002 is cubic in shape.22 The increased catalytic resolve its shortcomings, the combination of MOF-5 with GO
activity, conductivity, and surface area of MOFs needed to be is largely necessary. However, there are needs for MOF-5@GO
combined with other materials such as mesoporous silica/ nanocomposites such as one-step synthesis, ambient reaction
MOF, mesoporous alumina/MOF, and MOF/graphene have conditions, potential for catalyst recovery, recyclability, and
been reported for adsorption.23−29 Furthermore, carbon-based efficient and successful adsorbents for RhB in the treatment of
materials including activated carbon, CNT, and GO have also wastewater. This is a simple and solvent-free synthetic process
been incorporated with MOFs to improve the adsorptive to minimize environmental pollution. Herein, we report a one-
properties of composite materials.30 Owing to their remarkable step synthesis of MOF-5@GO nanocomposites employed for
chemical and physical characteristics, MOFs/GO-based nano- adsorption kinetic studies of RhB. Additionally, the effect of
composite materials are in tremendous demand for adsorption MOF-5 and GO ratio (1:1, 1:2, and 2:1) with pH, contact
of dye.31,32 In the last 10 years, the excellent and exclusive time, and concentration of RhB on the adsorption performance
chemical, electrical, and mechanical characteristics of graphene is also studied. Moreover, adsorption kinetics and adsorption
oxide have attracted remarkable interest.33 The process of isotherms are well described along with the plausible
chemical modification of graphene and graphene oxide mechanism of RhB adsorption onto MOF-5@GO nano-
includes covalent binding and non-covalent binding.34 The composites. The stability and self-degradation studies of RhB
structure of graphene is extremely stable, the bond between (up to 120 days) at different pH values (pH 1−12) are also
each carbon atom is strong, and the connection between reported.
carbon atoms via a covalent bond is strong. Graphene has a
simpler synthetic process and lower production cost as
2. RESULTS AND DISCUSSION
compared to other carbon-based materials. GO is a two-
dimensional material with hexagonal shape and has well- 2.1. FTIR and XRD Study. The FTIR and XRD spectra of
developed functional groups that contain oxygen moieties. pristine GO, MOF-5, MOF-5@GO, and RhB-MOF-5@GO
However, these oxygen-containing groups may increase nanocomposites are given in Figure 1. The FTIR spectrum of
catalytic activity, increase stability, and are also according to GO exhibits strong characteristics bands at 3414, 1734, 1618,
the requirements that need chemical modification.35 GO is 1224, 1041, and 866 cm−1 (Figure 1a(i)). The typical peaks
readily dispersed in water and some other polar solvents due to found in GO are the hydroxyl group of −OH stretching
its hydrophilic character, water stability, and solution vibrations at 3414 cm−1, the carbonyl or carboxylic moiety of
dispersibility.36 A series of MOF@GO composites of excellent CO stretching frequency at 1734 cm−1, the aromatic
adsorption capacity has been recorded by the Bandosz stretching frequency of CC present at 1618 cm−1, and the
group.24,37 However, the possible interactions between GO stretching vibration of C−O epoxy found at 1224 cm−1. The
and MOF-5 composites enhance the hydrophobic stability of C−O alkoxy stretching vibration appeared at 1041 cm−1, and
the MOF-5 framework.38,39 MOF-5 also has the characteristic the epoxy or peroxide functional group occurred at 866 cm−1
9588 https://doi.org/10.1021/acsomega.1c00143
ACS Omega 2021, 6, 9587−9599
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 2. High-resolution XPS of (a) C1s of GO, (b) C1s of MOF-5, (c) O1s of GO, (d) O1s of MOF-5, and (e) Zn2p metal. (f) Survey XPS
spectra of (i) GO, (ii) MOF-5, and (iii) MOF-5@GO nanocomposites.

of GO.42 The pure MOF-5 shows two strong bands present at The adsorption band slightly shifted, and the obtained bands
1657 and 1599 cm−1 corresponding to the CO and C−O are found at 2977, 1685, 1586, 1549, 1389, 1294, 932, and 757
stretching vibration of the aromatic ring of terephthalic acid. cm−1, which may be hydrogen bonding and confirm the RhB
The stretching frequency of C−H is found at 1387 cm−1, adsorbed on MOF-5@GO nanocomposites.44 The Zn−O
although the characteristics peaks were found at 1200−600 band is shifted and found to be 494 cm−1 due to interactions
cm−1 out-of-plane vibrations of an aromatic ring of terephthalic with water, GO, and RhB. However, after being immersed in
acid of the MOF-5 framework.43 The absorption band appears water up to 5 days, the MOF-5@GO nanocomposite did not
at 530 cm−1 can be assigned to the Zn−O stretching frequency lose its crystallinity. Additionally, some interactions from
as shown in Figure 1a(ii). FTIR spectra of MOF-5@GO oxygenated functional groups of graphene oxide with vacant
nanocomposites show that the majority of peaks for GO and sites of the zinc cluster of MOF-5 maintain the crystallinity as
MOF-5 are retained and there was no significant change. It was well as increase the conductivity of MOF-5 and enhanced the
also proven that GO did not affect the property of MOF-5, as adsorption properties, and this is consistent with a previous
given in Figure 1a(iii). FTIR spectra of RhB are shown in study.45 Furthermore, the FTIR spectrum of MOF-5@GO
Figure 1a(iv), and the absorption band displayed at 1683 cm−1 nanocomposites in water is observed up to the 7th day, which
can be assigned to the stretching frequency of CO. The is represented in Figure 1c. Subsequently, the interactions of
peaks at 2985 cm−1 indicated the stretching band of −CH. The the −OH group of the water molecule from vacant sites of the
peaks at 1586 cm−1 were attributed to the stretching vibration zinc cluster may create some disorderness of MOF-5 by
of the aromatic benzene ring of CC. The functional group of breaking of the Zn−O bond between the metal cluster and
RhB shows peaks at 1403 (−CH3), 1327 (CN), and 991 organic linker.
cm−1 for the C−H group. The RhB-adsorbed MOF-5@GO The XRD pattern of graphite, GO, MOF-5, MOF-5@GO,
nanocomposite is shown in Figure 1a(v). and RhB-MOF-5@GO nanocomposites is shown in Figure 1c.
9589 https://doi.org/10.1021/acsomega.1c00143
ACS Omega 2021, 6, 9587−9599
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 3. (a) (i) GO, (ii) MOF-5, (iii) RhB, and (iv) RhB-MOF-5@GO nanocomposites. (b) RL values.

Graphite shows a characteristic sharp peak at 2θ = 26.5° 2.3. Thermogravimetric Analysis. GO, MOF-5, and
attributed46 to ⟨002⟩ as shown in Figure 1c(i). Furthermore, RhB-MOF-5@GO nanocomposites were characterized by
GO shows typical peaks in Figure 1c(ii) for planes ⟨001⟩ and thermogravimetric analysis (TGA) and are shown in Figure
⟨100⟩ at 2θ = 10.3° and 46.3°, respectively. Pure MOF-5 has 3. GO shows four weight loss results between 25 and 600 °C as
shown crystalline peaks, which find 2θ values at 6.82, 9.65 shown in Figure 3a(i). The first weight loss observed between
13.65, 15.30, 20.59, 22.36, 24.71, 31.31, and 32.13 for crystal 25 and 106 °C (17.07%) can be assigned to the removal of
planes ⟨200⟩, ⟨220⟩, ⟨400⟩, ⟨420⟩, ⟨531⟩, ⟨533⟩, ⟨551⟩, ⟨751⟩, physically adsorbed water. The second weight loss between
and ⟨911⟩, respectively, as depicted in Figure 1c(iii). The 106 and 183 °C (9.39%) is attributed to the removal of labile
MOF-5@GO shows diffraction peaks detected at 2θ at 10.12, oxygen-containing functional groups. The third weight loss
15.61, 18.55, 20.58, 25.95, 27.27, 28.39, 30.31, 31.12, 32.34, between 183 and 220 °C (24.3%) indicates the decomposition
34.06, and 43.80, and these peaks are slightly shifted as of stable oxygen functionalities. The fourth weight loss
compared to pure MOF-5 but do not distort the crystalline between 220 and 600 °C (27.68%) is due to the full collapse
nature after the combination of GO as shown in Figure 1c(iv). of GO.
The RhB-adsorbed MOF-5@GO nanocomposite shows There are four weight loss results of MOF-5 depicted in
similar diffraction peaks for MOF-5@GO nanocomposites, Figure 3a(ii). The weight loss between 25 and 119 °C (2.67%)
which is shown in Figure 1c(v). The adsorbed molecule RhB is is due to the removal of physically adsorbed water. The weight
not crystalline, it is amorphous. Furthermore, XRD results loss between 119 and 215 °C (15.62%) is attributed to the
show that the crystalline nature of MOF-5@GO nano- removal of N,N-dimethylformamide. The weight loss between
composites did not distort up to the 5th day of MOF-5@ 215 and 330 °C (13.1%) is due to the partial decomposition of
GO nanocomposites47 as shown in Figure 1e. However, after 7 the MOF-5 frameworks. The weight loss between 330 and 524
days, the crystallinity of MOF-5@GO nanocomposites was lost °C (41.79%) is due to the collapse of the framework52 of
and might be some distorted after being immersed in water MOF-5 and constant up to 600 °C.
due to strong interactions from the free vacant d-orbital of the RhB shows three weight loss results between 176 and 600
zinc metal cluster,48,49 as given in Figure 1d. °C and stability up to 176 °C. The first weight loss between
2.2. X-ray Photoelectron Spectroscopy (XPS) Studies. 176 and 215 °C (4.07%) is attributed to the removal of water,
The synthesized GO, MOF-5, and MOF-5@GO nano- the second weight loss between 215 and 300 °C (8.89%) is
composites were further studied via XPS to define the credited to the removal of the oxygen functionality group, and
chemical states of different aspects and the presence of the third weight loss between 300 and 600 °C (65.22%) is
functional groups.50,51 The deconvoluted high-resolution XPS attributed to aromatic containing groups as illustrated in Figure
C1s region spectra of GO and MOF-5 are displayed in Figure 3a(iii).
2a,b, respectively. The C1s spectra of pristine GO have In the case of RhB-MOF-5@GO nanocomposites, there are
characteristic peaks at binding energies of 282.80, 284.78, five weight loss results. The first weight loss between 25 and 99
285.81, and 287.16 eV, representing C−C, C−O, and OC °C (7.32%) indicates the loss of physically adsorbed water.
O, and CO respectively (Figure 2a). Similarly, the high- The second weight loss between 97 and 167 °C (4.07%)
resolution XPS of the C1s region spectra of MOF-5 (Figure indicates the loss of oxygen functionality of the surface of GO
2b) have peaks with binding energies of 284.91, 286.0, and sheet. The third weight loss between 167 and 215 °C (12.76%)
288.86 eV, which can be attributed to C−C, CO, and O is due to the DMF molecule as well as the stable oxygen
CO, respectively. Figure 2c illustrates the XPS of O1s of GO functionality of GO and RhB functional groups, and the fourth
where peaks at binding energies of 529.55, 530.59, 531.34, and weight loss between 215 and 300 °C (18.28%) is attributed to
531.62 eV represent OCO, CO, C−OH, and C−O−C, the partial decomposition of the MOF-5 frameworks, GO, and
respectively. Moreover, the deconvoluted O1s of MOF-5 as RhB functional groups, while the fifth weight loss between 300
shown in Figure 2d depicts peaks at binding energies of 529.91 and 600 °C (23.57%) can be attributed to the full
and 531.49 eV representing OCO and CO, respec- decomposition of RhB-MOF-5@GO nanocomposites as
tively. Furthermore, peaks at binding energies of 1045.13 and shown in Figure 3a(iv).
1022.07 are attributed to Zn2p1/2 and Zn2p3/2 as represented 2.4. Surface Area Analysis. The N2 adsorption−
in Figure 2e. The full spectra of GO, MOF-5, and MOF-5@ desorption isotherm of GO (Figure S2a), MOF-5 (Figure
GO nanocomposites clearly showing the presence of carbon, S2b), and MOF-5@GO nanocomposites (Figure S2c) is
oxygen, and zinc metal are represented in Figure 2f. shown in the Supporting Information. The MOF-5@GO
9590 https://doi.org/10.1021/acsomega.1c00143
ACS Omega 2021, 6, 9587−9599
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 4. FESEM monograph of (a) GO, (b) MOF-5, and (c) RhB-MOF-5@GO. (d) HRTEM image of RhB-MOF-5@GO nanocomposites.

Figure 5. (a) Stability of RhB at pH 1−12. (b) Calibration plot of RhB.

nanocomposite of the surface area is found to be 135.37 m2 Figure 4a. The FESEM image of pure MOF-5 depicts a stacked
g−1, indicating the successful reaction with GO and MOF-5. elongated hexagonal structure as shown in Figure 4b. However,
The surface area of the MOF-5 is 29.95 m2 g−1, which is in the FESEM image in Figure 4c and HRTEM image in
greater than that from the literature, which is given as 12.35 m2 Figure 4d of RhB-MOF-5@GO nanocomposites, MOF-5
g−1.53 This value is too low due to air exposure during the formed an irregularly shaped structure of 50−200 nm instead
treatment, and the perfect BET surface area of MOF-5 as of an elongated hexagon, and this might be due to the facile
reported by the group of Yaghi54 reached up to 2900 m2 g−1. synthesis of the nanocomposite. Furthermore, energy-dis-
The surface area of GO is found to be 9.45 m2 g−1. The specific persive X-ray spectroscopy (EDS) of RhB-adsorbed MOF-
surface area of the synthesized nanocomposite material is 5@GO nanocomposites shows the elemental composition as
increased due to the combination of MOF-5 and GO. The well as mapping of elements such as C, O, N, Cl, and Zn as
calculated results indicate that the average pore size of the given in the Supporting Information, i.e., Figure S3a−g.
MOF-5@GO nanocomposite is found to be 0.014 nm. The 2.6. Influence of Process Variable on Adsorption.
higher surface area and pore volume of MOF-5@GO 2.6.1. Effect of pH on RhB. The absorbance spectra of the RhB
nanocomposites can afford extra possible sites for adsorption were recorded in an aqueous solution from the wavelength of
reactions. The MOF-5@GO nanocomposite displays a type III 350−650 nm in a quartz cell having a path length of 1 cm with
curve and H3 hysteresis loop, and this indicates that the MOF- λmax corresponding to RhB found at 554 nm. The absorbance
5@GO nanocomposite is shown as a mesoporous structure spectra of RhB show a pH-independent behavior at pH 4−12
with a relatively high surface area. and are shown in Figure 5a. There is no noticeable change that
2.5. Morphological Characterization. The morphologies occurs in this range. Thus, this confirms that RhB is more
of nanocomposite are studied by HRTEM and FESEM, and stable and does not lose its chemical behavior between pH 4
illustrations of the pristine GO sheet, pure MOF-5, and RhB- and 12.
adsorbed MOF-5@GO nanocomposites are shown in Figure 4. 2.6.2. Calibration Plot of RhB. The stock solution of RhB
The FESEM image of GO shows that it is a typical sheet-like was prepared in double-distilled water. A series of solutions
randomly aggregated thin crumpled layer structure as shown in with known concentrations (1−8 ppm) were prepared in a 25
9591 https://doi.org/10.1021/acsomega.1c00143
ACS Omega 2021, 6, 9587−9599
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 6. Removal of RhB at (a) 1:1, (b) 1:2, and (c) 2:1 ratios as a function of pH and (d) a function of time.

Figure 7. Percentage removal (a) effect of the initial concentration of RhB and (b) effect of Ni2+ and Cu2+ metal ions.

mL standard volumetric flask. The Lambert−Beer law is valid %removal of RhB


up to 6 ppm for RhB dye. RhB in aqueous solution shows a initial con. of RhB − final con. of RhB
= × 100
pH-independent absorption behavior at a range of 4−12. initial con. of RhB (1)
However, the calibration plot of RhB was prepared at pH 7,
2.6.4. Removal of RhB as a Function of pH. In the
and the correlation coefficient (R2) was found to be 0.996 and
adsorption method, pH plays an important role in RhB. This
is shown in Figure 5b. Hence, at pH 7, further estimation of study was done at 50 mg of the nanocomposite in different
RhB has been carried out. ratios of MOF-5 and GO and 25 mL of 20 ppm RhB at
2.6.3. Batch Extraction Study and Removal of RhB. Using different pH values like 2, 4, 5, 6, 8, and 12 at room
the batch extraction process, the adsorption behavior of RhB temperature as given in Figure 6. Adsorption of RhB on the
on MOF-5@GO nanocomposites was studied as a function of ratio of MOF-5 and GO such as 1:1, 1:2, and 2:1 is shown in
pH, contact time, and initial concentration of the aqueous RhB Figures 66b, and 6ca, , respectively. The highest removal of
solution. Each experiment was done with 50 mg of the RhB was observed on the ratio of 1:1 in comparison to 1:2 and
2:1. Henceforth, the 1:1 ratio is selected for further batch
nanocomposite at room temperature. The unabsorbed RhB
study. The amount of RhB adsorbed on MOF-5@GO
concentration present in the supernatant was calculated by nanocomposites was found in a higher acidic medium (at
UV−Vis spectrophotometry. The percentage (%) removal of pH 2) as compared to the basic medium and remained almost
RhB on the MOF-5@GO nanocomposite was determined constant from pH 6 to 8. So, consequently, pH 2 and 1:1 ratio
using eq 1: are chosen for further study.
9592 https://doi.org/10.1021/acsomega.1c00143
ACS Omega 2021, 6, 9587−9599
ACS Omega http://pubs.acs.org/journal/acsodf Article

Table 1. Parameters of the Kinetic and Adsorption Isotherm


pseudo-first-order kinetics pseudo-second-order kinetics
K1 (min−1) qe (mg/g) R2 K2 [g/mg·min)] qe (mg/g) R2
0.4470 9.6272 0.4025 0.0367 26.7379 0.9908
Freundlich adsorption isotherm Langmuir adsorption isotherm
1/n n Kf R2 qmax (mg/g) KL (dm3/mg) R2 RL
0.5254 1.9033 9.4362 0.0861 169.4915 0.0302 0.9703 0.1418

2.6.5. Removal of RhB as a Function of Contact Time. In Lagergren in which the plot between log qe−qt against t gives a
the study, 50 mg of MOF-5@GO nanocomposites was used straight line as shown in Figure S1a, pseudo-second-order
for 25 mL of 20 ppm RhB for adsorption while maintaining the kinetics is described by Ho and McKay in which the plot
pH at 2. Approximately 98.88% of RhB is removed within 5 between t/qt against t gives a straight line (Figure S1b), and
min of contact time, which increased to 99.68% within 10 min, the intraparticle diffusion (ID) model given by Weber and
and this may be due to the availability of more adsorption sites Morris in which the plot between qt against gives a straight line
due to a high surface area. However, adsorption remained (Figure S1c) in which the correlation coefficient R2 is found to
almost constant after 15 min as shown in Figure 6d. The be 0.7779. The experimental parameters of pseudo-first-order
reason thereof, during the initial few minutes, is that the rate of and pseudo-second-order kinetics are listed in Table 1. From
removal of RhB is very fast due to the concentration gradient the experimental parameter based on the correlation coefficient
that formed at the start of the process between RhB and the value, it is observed that pseudo-second-order kinetics is best
adsorbent surface. This effect is clarified by the presence of fitted as compared to pseudo-first-order kinetics.
vacant sites on the surface of MOF-5@GO nanocomposites The adsorption capacity of MOF-5@GO nanocomposites
available for adsorption. Furthermore, due to the saturation of for RhB is also determined by the equilibrium adsorption
adsorption sites or blockage of pore on MOF-5@GO study. There are three types of adsorption isotherm. Langmuir
nanocomposites, the adsorption percent of RhB becomes adsorption isotherm defines the development of the monolayer
constant. surface of RhB quantitatively around the adsorbent in which
2.6.6. Effect of the Initial Concentration of RhB. The the plot between Ce/qe against Ce gives a straight line as shown
removal of RhB by MOF-5@GO nanocomposites is also in Figure S1d, and it can also be defined in terms of a
affected by the initial amount of RhB present in the aqueous dimensionless constant known as Hall separation factor (RL)
solution. Therefore, the consequence of initial concentration shown in Figure 3b, which is found to be 0.1418. Freundlich
on the removal of RhB is investigated with 25 mL of RhB with adsorption isotherm explains the adsorption process in the
50 mg of MOF-5@GO nanocomposites at pH 2, and the heterogeneous systems and the formation of multilayer
concentration of RhB that varies from 5 to 550 ppm is shown
adsorption in which the plot between log qe against log Ce
in Figure 7a. The rate of removal of RhB by MOF-5@GO
gives a straight line as shown in Figure S1e. Temkin adsorption
nanocomposites increases up to 500 ppm; however, due to
isotherm explains the heat of adsorption of the entire molecule
saturation of available loading sites, no significant increment in
in the adsorbent−adsorbate interaction in which the plot
adsorption percentage of RhB is further observed.
between qe against Ce gives a straight line as shown in Figure
The removal of RhB was found to be 60.64% (151.62 mg·
g−1) at 500 ppm; therefore, RhB adsorbed on MOF-5@GO S1f.
nanocomposites. Hence, to investigate the structure and Experimental parameters of the Langmuir and Freundlich
morphology of MOF-5@GO nanocomposite with appropriate adsorption isotherm are listed in Table 1. From the
analytical techniques, it was prepared in large amount. experimental parameter based on the correlation coefficient
2.6.7. Effect of Metal Concentration on RhB. The effect of value, it is observed that the Langmuir adsorption isotherm is
metal ion concentration also played a crucial role in the best fitted as compared to the Freundlich adsorption isotherm
remediation of RhB. The effect of metal ions Ni2+ and Cu2+ at in the adsorption of RhB on to the MOF-5@GO nano-
different concentrations (5−100 ppm) on the removal of RhB composite.
on MOF-5@GO nanocomposites has been studied. In the 2.8. Natural Degradation of RhB Concerning pH. Self-
presence of Cu2+, the percentage removal efficiency of RhB on degradation studies of RhB were done to see the degradation
MOF-5@GO nanocomposites decreases with an increase in efficiency process naturally at different pH values (pH 1−12)
the concentration of a metal ion, but in the case of Ni2+ ion, at a time duration of 1−120 days and are shown in Figure 8a−
the percentage removal of RhB on MOF-5@GO nano- l. In acidic conditions, at pH 1, 12.40% (Figure 8a), there is a
composites slightly decreases with increasing the concentration small degradation that takes place, but in pH 2, 35.87% (Figure
of a metal ion in aqueous solution.55 In the presence of heavy 8b) degraded within 60 days and it remains constant up to 120
metals, the adsorption capacity of dyes onto the MOF-5@GO days. In pH 3, there is no degradation up to 60 days but
surface decreases due to the preferential adsorption of these 21.59% (Figure 8c) degradation within 90 days takes place and
metal ions onto the active site of the MOF-5@GO it remains constant up to 120 days, but in the case of pH 4 and
nanocomposite shown in Figure 7b. pH 5, there is no degradation up to 90 and 30 days,
2.7. Description of Adsorption Kinetics and Iso- respectively, but in both cases, 30.25% (Figure 8d) and 21.72%
therms. Chemical kinetics deals with the rate of reaction, as (Figure 8e) degraded up to 120 days. In the case of pH 6, there
well as its mechanism, which is given in the Supporting are slight decreases after 30 days (Figure 8f). From pH 7 to pH
Information. In the adsorption process, there are three types of 11, there is no degradation of RhB that was found up to 120
adsorption kinetics. Pseudo-first-order kinetics is defined by days (Figure 8g−k), but in the case of the strong basic medium
9593 https://doi.org/10.1021/acsomega.1c00143
ACS Omega 2021, 6, 9587−9599
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 8. Degradation of RhB for 1−120 days at various pH values: (a) pH 1, (b) pH 2, (c) pH 3, (d) pH 4, (e) pH 5, (f) pH 6, (g) pH 7, (h) pH
8, (i), pH 9, (j) pH 10, (k) pH 11, and (l) pH 12.

According to the available reports,56,57 the pH of the


at pH 12, slight degradation was found, which is 36.16% seawater, river, sewage, and lake water may exist from pH 6 to
(Figure 8l) from 15 to 120 days. pH 8, and from experimental data, it is evidenced that RhB
9594 https://doi.org/10.1021/acsomega.1c00143
ACS Omega 2021, 6, 9587−9599
ACS Omega http://pubs.acs.org/journal/acsodf Article

Figure 9. (a) Removal efficiency of RhB on MOF-5@GO nanocomposites in different cycles. (b) SEM image of MOF-5@GO nanocomposites
after the fifth run.

does not undergo any degradation in this region. It can be cationic dye with highly efficient adsorbents MOF-5@GO
concluded that the RhB exists for a long time in an aqueous nanocomposites.
medium and it is imperative to remediate RhB.
2.9. Adsorption Mechanism of RhB on MOF-5@GO Table 2. Adsorption Capacity of RhB onto MOF-5@GO
Nanocomposites. The adsorption of RhB on the surface and Nanocomposites as Compared with Previous Studiesa
the cavity of MOF-5@GO nanocomposites can be due to
different factors such as electrostatic/ionic interactions, Qe conditions
adsorption adsorbate (mg/g) (°C) ref
hydrogen bonding, and π−π interactions.58−60 Keeping this
in mind, it can be proposed that electrostatic attraction takes CoOF RhB 72.15 20 °C 63
place between positively charged RhB molecules and RGO RhB 13.15 25 °C 64
negatively charged −OH groups of the surface of the MOF- sodium montmorillonite RhB 42.19 30 °C 65
5@GO nanocomposites. The MOF-5@GO nanocomposite CNSs RhB 01.95 25 °C 66
has great advantages, such as high surface area, variable pore MCM-22 RhB 01.05 30 °C 67
sizes, and different functionalities. Figure 1a(v) shows the Perlite RhB 08.72 30 °C 68
FTIR spectrum for the comparison of MOF-5@GO nano- Nano-NiO RhB 111.0 20 °C 69
composite before and after adsorption of RhB. The shifted MOF-5@GO RhB 151.5 RT PW
nanocomposite
peak position is because of the hydrogen bonding, π−π a
PW, present work; RT, room temperature.
interactions, and electrostatic attraction. The removal of dye is
usually based on the effect of size and the importance of ionic
selectivity. Lan et al. created a mesoporous MOF of about 38 Å 2.11. Cyclability. The recycling ability of MOF-5@GO
size-tunable cages and investigated that the size-exclusion nanocomposites is also established where it is renewed by
effect depends on the separation of large dye molecules.61 The centrifugation after the batch experiment followed by excessive
widely accepted theories accounting for the adsorption washing with ethanol and dried at 75 °C before further use in
mechanism of adsorbents are based on their ionic selectivity the next cycle. The recycling performance of regenerated
and size effect. It may also be possible that the cause is due to MOF-5@GO nanocomposites is shown in Figure 10a. An
the filling of pores occurring in the MOF-5 and GO, which is insignificant decrease in adsorption percentage is observed
the inhibition of the adsorption of RhB to that of the after the first three cycles, however remaining constant for the
synthesized MOF-5@GO nanocomposites. Including the high last two cycles.
pore volume also allows them to absorb a greater amount of This slight decrement in adsorption efficiency of nano-
RhB in an aqueous solution. The mechanism for the composites might be due to the movement of the RhB
adsorption of RhB on MOF-5@GO nanocomposites can be molecule toward the interior cavities of the MOF-5@GO
ionic interactions due to presence of unsaturated bonds (C nanocomposite and it could not be washed out easily. The
C, O−CO) in MOF-5, negatively charged GO surface, and SEM image of MOF-5@GO nanocomposites depicted in
positive charge of RhB.62 Furthermore, π−π interaction also Figure 10b after the fifth cycle also shows that there are no
plays as active binding sites for the adsorption of the RhB significant changes occurring in the morphology. These results
molecule with MOF-5@GO nanocomposites. Since RhB demonstrated quicker adsorption integrated with excellent
contains CC double bonds and π-electrons, this π-electron reversibility, thus making it a promising candidate for RhB
can easily interact with the π-electrons of benzene rings on the adsorption in real applications. Furthermore, as compared with
MOF-5 and GO surface via π−π interactions. Based on the previous studies, the MOF-5@GO nanocomposite displays the
above facts of interactions, RhB easily got adsorbed on the highest adsorption capacity, which is listed in Table 2.
surface of MOF-5@GO nanocomposites. A schematic
representing the proposed mechanism of RhB adsorption is 3. CONCLUSIONS
shown in Figure 9. We have successfully developed MOF-5@GO nanocomposites
2.10. Comparison Study. RhB removal capacities by the employed for the removal of toxic dye. Due to electrostatic
synthesized nanocomposites showing greater effectiveness as interactions between the positively charged metal center of
compared with previously reported studies are shown in Table MOF-5 and negative charge on GO surface, MOF-5 and GO
2. The synthesized nanocomposites demonstrate a strong resulted in MOF-5@GO nanocomposite. MOF-5@GO nano-
9595 https://doi.org/10.1021/acsomega.1c00143
ACS Omega 2021, 6, 9587−9599
ACS Omega http://pubs.acs.org/journal/acsodf Article

4. EXPERIMENTAL SECTION
4.1. Synthesis of Graphene Oxide. GO was synthesized
according to previous reports,70−72 with some modifications
via the improved Hummer’s method. In brief, 90:10 mL of
concentrated H2SO4:H3PO4 and graphite powder (1.5 g) were
added with constant stirring for 15 min. Afterward, potassium
permanganate (8 g) was added slowly into the above mixture
and was heated at 50−55 °C with constant stirring for 12 h.
The reaction was cooled to room temperature and poured into
ice-cold water followed by the addition of hydrogen peroxide
(3.5 mL) and stirred for 1 h, yielding the bright yellow
precipitate. The mixture was kept overnight and centrifuged for
20 min at 7000 rpm (Sigma, Laboratory Centrifuges 3K30
Sartorius). The remaining solid material was then washed with
hydrogen chloride, ethanol (2x), and diethyl ether (2x) twice
for purification. The final solid brownish color GO sample was
Figure 10. Schematic possible interaction in the adsorption vacuum dried at 45 °C for 10 h.
mechanism of RhB on MOF-5@GO nanocomposites.
4.2. Synthesis of MOF-5 and MOF-5@GO Nano-
composite. Zinc acetate (4.24 g) and benzene dicarboxylic
composites have an excellent adsorption capacity at a ratio of acid (1.26 g) were dissolved in N,N-dimethylformamide
1:1 for RhB, proving its application in the remediation of toxic (DMF) in a beaker with constant stirring up to 2 h to obtain
dye. The physicochemical properties of nanocomposites were a homogeneous mixture and placed in an oven at 125 °C for
explained with different structural and morphological charac- 24 h in a Teflon-lined sealed solvothermal vessel, resulting in a
terization techniques. The morphology of RhB-adsorbed solid colorless precipitate. It was gradually cooled to room
MOF-5@GO nanocomposites reveals the growth of irregular temperature and centrifuged at 7000 rpm for 15 min. The
form MOF-5 nanoparticles with sizes varying between 50 and product was washed with N,N-dimethylformamide (100 x 2),
200 nm. Results of XPS and EDS spectra revealed the presence dried under vacuum at room temperature, and stored under a
of carbon, oxygen, and zinc metal cluster. The adsorption desiccator.
kinetics data reveals pseudo-second-order (R2 = 0.9908), MOF-5 was synthesized according to a previously reported
although the adsorption isotherm model was well defined by method73,74 with minor modifications. MOF-5@GO nano-
the Langmuir adsorption isotherm (R2 = 0.9703). The effect of composites were synthesized by a grinding method at different
adsorption capacity in the presence of heavy metal ions, i.e., ratios (1:1, 1:2, and 2:1). The mixture was then dried at room
Ni2+ and Cu2+, shows a decrease in the presence of Cu2+ ion temperature for 12 h and used as an adsorbent, and the
but insignificant in the presence of Ni2+ ion. The adsorption synthesis is schematically depicted in Scheme 1.


capacity of MOF-5@GO nanocomposites was found to be
99.58% with 20 ppm at pH 2. In conclusion, MOF-5@GO ASSOCIATED CONTENT
nanocomposite is synthesized in one step, it is environmentally
friendly, and cost-effective with high adsorption efficiency and *
sı Supporting Information

outstanding reusability, which makes it a better adsorbent The Supporting Information is available free of charge at
material for the removal of RhB from waste water. https://pubs.acs.org/doi/10.1021/acsomega.1c00143.

Scheme 1. Simple Representation of the Synthesis of MOF-5@GO Nanocomposites

9596 https://doi.org/10.1021/acsomega.1c00143
ACS Omega 2021, 6, 9587−9599
ACS Omega http://pubs.acs.org/journal/acsodf Article

Materials and method; description of adsorption kinetics synthetic dyes by the laccase from a soil-isolated ascomycete,
and adsorption isotherm; N2 adsorption−desorption of Paraconiothyrium variabile. Int. Biodeterior. Biodegrad. 2013, 85,
GO, MOF-5, and MOF-5@GO nanocomposites; and 173−181.
EDS elemental analysis (PDF) (11) Liu, Y.; Ng, Z.; Khan, E. A.; Jeong, H. K.; Ching, C. B.; Lai, Z.


Synthesis of continuous MOF-5 membranes on porous a-alumina
substrates. Microporous Mesoporous Mater. 2009, 118, 296−301.
AUTHOR INFORMATION (12) Dalvand, A.; Nabizadeh, R.; Ganjali, M. R.; Khoobi, M.;
Corresponding Author Nazmara, S.; Mahvi, A. H. Modeling of Reactive Blue 19 azo dye
removal from colored textile wastewater using L-arginine-function-
Dhanraj T. Masram − Department of Chemistry, University of
alized Fe3O4 nanoparticles: Optimization, reusability, kinetic and
Delhi, Delhi 110007, India; orcid.org/0000-0001-6273-
equilibrium studies. J. Magn. Magn. Mater. 2016, 404, 179−189.
8199; Email: dhanraj_masram27@rediffmail.com (13) Mirzadeh, S. S.; Khezri, S. M.; Rezaei, S.; Forootanfar, H.;
Author Mahvi, A. H.; Faramarzi, M. A. Decolorization of two synthetic dyes
using the purified laccase of Paraconiothyrium variabile immobilized
Gyanendra Kumar − Department of Chemistry, University of
on porous silica beads. J. Environ. Health Sci. Eng. 2014, 12, 1−9.
Delhi, Delhi 110007, India; orcid.org/0000-0002-1975- (14) Bazrafshan, E.; Mostafapour, F. K.; Hosseini, A. R.; Khorshid,
5310 A. R.; Mahvi, A. H. Decolorisation of Reactive Red 120 Dye by Using
Complete contact information is available at: Single-Walled Carbon Nanotubes in Aqueous Solutions. Journal of
https://pubs.acs.org/10.1021/acsomega.1c00143 Chemistry 2013, 2013, 1−8.
(15) Sivakumar, P.; Palanisamy, P. N. Adsorption studies of basic
Notes red 29 by a nonconventional activated carbon prepared from
The authors declare no competing financial interest. Euphorbia antiquorum L. Int. J. ChemTech Res. 2009, 1, 502−510.


(16) Mogha, N. K.; Gosain, S.; Masram, D. T. Gold nanoworms
immobilized graphene oxide polymer brush nanohybrid for catalytic
ACKNOWLEDGMENTS degradation studies of organic dyes. Appl. Surf. Sci. 2017, 396, 1427−
We are grateful to the Head, Department of Chemistry, and 1434.
Director, USIC, University of Delhi, India for providing (17) Dreyer, D. R.; Park, S.; Bielawski, C. W.; Ruoff, R. S. The
instrumentation facilities. The authors acknowledge SAIF- chemistry of graphene oxide. Chem. Soc. Rev. 2010, 39, 228−240.
AIIMS, New Delhi, for HRTEM facilities. G.K. is also thankful (18) Zhu, J.; Yang, D.; Yin, Z.; Yan, Q.; Zhang, H. Graphene and
to the UGC-SRF (file no. 19/06/2016(i)EU-V) for providing graphene-based materials for energy storage applications. Small 2014,
financial support. 10, 3480−3498.


(19) Janiak, C.; Vieth, J. K. MOFs, MILs and more: concepts,
properties and applications for porous coordination networks
REFERENCES (PCNs). New J. Chem. 2010, 34, 2366−2388.
(1) Ai, L.; Zhang, C.; Meng, L. Adsorption of Methyl Orange from (20) Wang, Y.; Hou, C.; Zhang, Y.; He, F.; Liu, M.; Li, X.
Aqueous Solution on Hydrothermal Synthesized Mg-Al Layered Preparation of graphene nano-sheet bonded PDA/MOF micro-
Double Hydroxide. J. Chem. Eng. Data 2011, 56, 4217−4225. capsules with immobilized glucose oxidase as a mimetic multi-enzyme
(2) Li, Z.; Ma, B.; Zhang, X.; Sang, Y.; Liu, H. One-pot synthesis of system for electrochemical sensing of glucose. J. Mater. Chem. B 2016,
BiOCl nanosheets with dual functional carbon for ultra-highly 4, 3695−3702.
efficient photocatalytic degradation of RhB. Environ. Res. 2020, 182, (21) Fan, W.; Wang, X.; Xu, B.; Wang, Y.; Liu, D.; Zhang, M.;
109077. Shang, Y.; Dai, F.; Zhang, L.; Sun, D. Amino-functionalized MOFs
(3) Wang, J.; Qin, L.; Lin, J.; Zhu, J.; Zhang, Y.; Liu, J.; van der with high physicochemical stability for efficient gas storage/
Bruggen, B. Enzymatic construction of antibacterial ultrathin separation, dye adsorption and catalytic performance. J. Mater.
membranes for dyes removal. Chem. Eng. J. 2017, 323, 56−63. Chem. A 2018, 6, 24486.
(4) Liu, E.; Du, Y.; Bai, X.; Fan, J.; Hu, X. Synergistic improvement (22) Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.;
of Cr (VI) reduction and RhB degradation using RP/g-C3N4
O’Keeffe, M.; Yaghi, O. M. Systematic design of pore size and
photocatalyst under visible light irradiation. Arabian Journal of
functionality in isoreticular MOFs and their application in methane
Chemistry 2020, 13, 3836−3848.
storage. Science 2002, 295, 469−472.
(5) Idris, M. N.; Ahmad, Z. A.; Ahmad, M. A. Adsorption
(23) Kumar, G.; Mogha, N. K.; Masram, D. T. Zr-Based Metal-
equilibrium of malachite green dye onto rubber seed coat based
activated carbon. Int. J. Basic Appl. Sci. 2011, 11, 38−43. Organic Framework/Reduced Graphene Oxide Composites for
(6) Liu, L.; Gao, Z. Y.; Su, X. P.; Chen, X.; Jiang, L.; Yao, J. M. Catalytic Synthesis of 2,3-Dihydroquinazolin4(1H)-one Derivatives.
Adsorption Removal of Dyes from Single and Binary Solutions Using ACS Appl. Nano Mater. 2021, DOI: 10.1021/acsanm.0c03322.
a Cellulose-based Bioadsorbent. ACS Sustainable Chem. Eng. 2015, 3, (24) Petit, C.; Bandosz, T. J. Exploring the coordination chemistry
432−442. of MOF-graphite oxide composites and their applications as
(7) Menzel, R.; Iruretagoyena, D.; Wang, Y.; Bawaked, S. M.; adsorbents. Dalton Trans. 2012, 41, 4027.
Mokhtar, M.; Al-Thabaiti, S. A.; Basahel, S. N.; Shaffer, M. S. P. (25) Bandosz, T. J.; Petit, C. MOF/graphite oxide hybrid materials:
Graphene oxide/mixed metal oxide hybrid materials for enhanced exploring the new concept of adsorbents and catalysts. Adsorption
adsorption desulfurization of liquid hydrocarbon fuels. Fuel 2016, 181, 2010, 17, 5−16.
531−536. (26) Pu, S.; Zhu, R.; Ma, H.; Deng, D.; Pei, X.; Qi, F.; Chu, W.
(8) Wong, Y. C.; Szeto, Y. S.; Cheung, W. H.; Mckay, G. Equilibrium Facile in-situ design strategy to disperse TiO2 nanoparticles on
Studies for Acid Dye Adsorption onto Chitosan. Langmuir 2003, 19, graphene for the enhanced photocatalytic degradation of rhodamine
7888. 6G. Appl. Catal. B 2017, 218, 208−219.
(9) Mariyam, A.; Shahid, M.; Mantasha, I.; Khan, M. S.; Ahmad, M. (27) MiarAlipour, S.; Friedmann, D.; Scott, J.; Amal, R. TiO2/
S. Tetrazole Based Porous Metal Organic Framework (MOF): porous adsorbents: Recent advances and novel applications. J. Hazard.
Topological Analysis and Dye Adsorption Properties. J. Inorg. Mater. 2018, 341, 404−423.
Organomet. Polym. 2020, 30, 1935−1943. (28) Ramesha, G. K.; Kumar, A. V.; Muralidhara, H. B.; Sampath, S.
(10) Ashrafi, S. D.; Rezaei, S.; Forootanfar, H.; Mahvi, A. H.; Graphene and graphene oxide as effective adsorbents toward anionic
Faramarzi, M. A. The enzymatic decolorization and detoxification of and cationic dyes. J. Colloid Interface Sci. 2011, 361, 270−277.

9597 https://doi.org/10.1021/acsomega.1c00143
ACS Omega 2021, 6, 9587−9599
ACS Omega http://pubs.acs.org/journal/acsodf Article

(29) Liu, T.; Li, Y.; Du, Q.; Sun, J.; Jiao, Y.; Yang, G.; Wang, Z.; Xia, highly stable MOF-5@MWCNTs nanocomposite with improved
Y.; Zhang, W.; Wang, K.; Zhu, H.; Wu, D. Adsorption of methylene hydrophobic properties. Arabian J. Chem. 2018, 11, 26−33.
blue from aqueous solution by graphene. Colloids Surf., B 2012, 90, (48) Greathouse, J. A.; Allendorf, M. D. The interaction of water
197−203. with MOF-5 simulated by molecular dynamics. J. Am. Chem. Soc.
(30) Petit, C.; Bandosz, T. J. MOF-graphite oxide nanocomposites: 2006, 128, 10678−10679.
surface characterization and evaluation as adsorbents of ammonia. J. (49) Han, S. S.; Choi, S.-H.; van Duin, A. C. T. Molecular dynamics
Mater. Chem. 2009, 19, 6521. simulations of stability of metal-organic frameworks against H2O
(31) Gao, W.; Alemany, L. B.; Ci, L.; Ajayan, P. M. New insights using the ReaxFF reactive force field. Chem. Commun. 2010, 46,
into the structure and reduction of graphite oxide. Nature Chem. 5713−5715.
2009, 1, 403−408. (50) Al-Gaashani, R.; Najjar, A.; Zakaria, Y.; Mansour, S.; Atieh, M.
(32) Compton, O. C.; Nguyen, S. T. Graphene oxide, highly reduced A. XPS and structural studies of high-quality graphene oxide and
graphene oxide, and graphene: versatile building blocks for carbon- reduced graphene oxide prepared by different chemical oxidation
based materials. Small 2010, 6, 711−723. methods. Ceram. Int. 2019, 45, 14439−14448.
(33) Dreyer, D. R.; Todd, A. D.; Bielawski, C. W. Harnessing the (51) Chul, H. D.; Vinodh, R.; Gopi, C. V. V. M.; Deviprasath, C.;
chemistry of graphene oxide. Chem. Soc. Rev. 2009, 43, 228−240. Kim, H.-J.; Yi, M. Effect of the cobalt and zinc ratio on the
(34) Li, Q.; Fan, F.; Wang, Y.; Feng, W.; Ji, P. Enzyme preparation of zeolitic imidazole frameworks (ZIFs): synthesis,
Immobilization on Carboxyl-Functionalized Graphene Oxide for characterization and supercapacitor applications. Dalton Trans.
Catalysis in Organic Solvent. Ind. Eng. Chem. Res. 2013, 52, 6343− 2019, 48, 14808−14819.
6348. (52) Yuan, J.; Liu, W.; Zhang, X.; Zhang, Y.; Yang, W.; Lai, W.; Li,
(35) Liao, Y.; Gao, Y.; Zhu, S.; Zheng, J.; Chen, Z.; Yin, C.; Lou, X.; X.; Zhang, J.; Li, X. MOF derived ZnSe-FeSe2/RGO Nano-
Zhang, D. Facile Fabrication of N-doped Graphene as Efficient composites with enhanced sodium/potassium storage. J. Power
Electrocatalyst for Oxygen Reduction Reaction. ACS Appl. Mater. Sources 2020, 455, 227937.
Interfaces 2015, 7, 19619−19625. (53) Lestari, W. W.; Wibowo, A. H.; Astuti, S.; Irwinsyah
(36) Park, S.; An, J.; Jung, I.; Piner, R. D.; An, S. J.; Li, X.; Pamungkas, A. Z.; Krisnandi, Y. K. Fabrication of hybrid coating
Velamakanni, A.; Ruoff, R. S. Colloidal suspensions of highly reduced material of polypropylene itaconate containing MOF-5 for CO2
graphene oxide in a wide variety of organic solvents. Nano Lett. 2009, capture. Prog. Org. Coat. 2018, 115, 49−55.
9, 1593−1597. (54) Li, H.; Eddaoudi, M.; O’Keeffe, M.; Yaghi, O. M. Design and
(37) Petit, C.; Bandosz, T. J. Enhanced adsorption of ammonia on synthesis of an exceptionally stable and highly porous metal-organic
metal-organic framework/graphite oxide composites: analysis of framework. Nature 1999, 402, 276−279.
surface interactions. Adv. Funct. Mater. 2010, 20, 111e118. (55) Kant, A.; Datta, M. Adsorption characteristics of Victoria blue
(38) Firouzjaei, M. D.; Shamsabadi, A. A.; Aktij, S. A.; Seyedfour, S. on low-cost natural sand and its removal from aqueous media. Eur.
F.; Gh, M. S.; Rahimpour, A.; Esfahani, M. R.; Ulbricht, M.; Soroush, Chem. Bull 2014, 3, 752−759.
M. Exploiting Synergetic Effects of Graphene Oxide and a Silver- (56) Feng, Z.; Su, B.; Xiao, D. D.; Ye, L. Y. Study on pH value and
Based Metal-Organic Framework to Enhance Antifouling and Anti- its variation characteristics of the main rivers into Dianchi lake under
Biofouling Properties of Thin-Film Nanocomposite Membranes. ACS the anthropogenic and natural processes. Yunnan, China, Taylor and
Appl. Mater. Interfaces 2018, 10, 42967−42978. Francis 2017, 38, 1197−1210.
(39) Yeh, C.-N.; Raidongia, K.; Shao, J.; Yang, Q.-H.; Huang, J. On (57) Zhang, C.; He, D.; Ma, J.; Tang, W.; Waite, T. D. aradaic
the origin of the stability of graphene oxide membranes in water. Nat. reactions in capacitive deionization (CDI) - problems and
Chem. 2015, 7, 166−170. possibilities: A review. Water Res. 2018, 128, 314−330.
(40) Petit, C.; Bandosz, T. J. MOF-Graphite Oxide Composites: (58) Yua, Y.; Murthya, B. N.; Shaptera, J. G.; Constantopoulosa, K.
Combining the Uniqueness of Graphene Layers and Metal-Organic T.; Voelckerb, N. H.; Ellis, A. V. Benzene carboxylic acid derivatized
Frameworks. Adv. Mater. 2009, 21, 4753−4757. graphene oxide nanosheets on natural zeolites as effective adsorbents
(41) Petit, C.; Burress, J.; Bandosz, T. J. The synthesis and for cationic dye removal. J. Hazard. Mater. 2013, 260, 330−338.
characterization of copper-based Metal-organic framework/graphite (59) Mittal, H.; Mishra, S. B. Gum ghatti and Fe3O4 magnetic
oxide composites. Carbon 2011, 49, 563−572. nanoparticles-based nanocomposites for the effective adsorption of
(42) Liu, L.; Zhang, B.; Zhang, Y.; He, Y.; Huang, L.; Tan, S.; Cai, X. rhodamine B. Carbohydr. Polym. 2014, 101, 1255−1264.
Simultaneous Removal of Cationic and Anionic Dyes from Environ- (60) Yang, C.; Wu, S.; Cheng, J.; Chen, Y. Indium-based metal-
mental Water Using Montmorillonite-Pillared Graphene Oxide. J. organic framework/graphite oxide composite as an efficient adsorbent
Chem. Eng. Data 2015, 60, 1270−1278. in the adsorption of rhodamine B from aqueous solution. J. Alloys
(43) Karimzadeh, Z.; Javanbakht, S.; Namazi, H. Carboxymethyl- Compd. 2016, 687, 804−812.
cellulose/MOF-5/Graphene oxide bio-nanocomposite as antibacterial (61) Lan, Y. Q.; Jiang, H. L.; Li, S. L.; Xu, Q. Mesoporous metal-
drug nanocarrier agent. Bioimpacts 2018, 9, 5−13. organic frameworks with size-tunable cages: selective CO2 uptake,
(44) Lamdab, U.; Wetchakun, K.; Kangwansupamonkon, W.; encapsulation of Ln3+ cations for luminescence, and column-
Wetchakun, N. Effect of a pH-controlled co-precipitation process chromatographic dye separation. Adv. Mater. 2011, 23, 5015−5020.
on rhodamine B adsorption of MnFe2O4 nanoparticles. RSC Adv. (62) Haque, E.; Jun, J. W.; Jhung, S. H. Adsorptive removal of
2018, 8, 6709. methyl orange and methylene blue from aqueous solution with a
(45) Jabbari, V.; Veleta, J. M.; Zarei-Chaleshtori, M.; Gardea- metal-organic framework material, iron terephthalate (MOF-235). J.
Torresdey, J.; Villagran, D. Green synthesis of magnetic MOF@GO Hazard. Mater. 2011, 185, 507−511.
and MOF@CNT hybrid nanocomposites with high adsorption (63) Barylak, M.; Cendrowski, K.; Mijowska, E. Application of
capacity towards organic pollutants. Chem. Eng. J. 2016, 304, 774− Carbonized Metal-Organic Framework as Efficient Adsorbent of
783. Cationic Dye. Ind. Eng. Chem. Res. 2018, 57, 4867−4879.
(46) Divya, K. S.; Chandran, A.; Reethu, V. N.; Mathew, S. (64) Sun, H.; Cao, L.; Lu, L. Magnetite/reduced graphene oxide
Enhanced photocatalytic performance of RGO/Ag nanocomposites nanocomposites: One step solvothermal synthesis and use as a novel
produced via a facile microwave irradiation for the degradation of platform for removal of dye pollutants. Nano Res. 2011, 4, 550−562.
Rhodamine B in aqueous solution. Appl. Surf. Sci. 2018, 444, 811− (65) Selvam, P. P.; Preethi, S.; Basakaralingam, P.; Thinakaran, N.;
818. Sivasamy, S.; Sivanesan, S. Removal of rhodamine B from aqueous
(47) Rehman, A. U.; Tirmizi, S. A.; Badshah, A.; Ammad, H. M.; solution by adsorption onto sodium montmorillonite. J. Hazard.
Jawad, M.; Abbas, S. M.; Rana, U. A.; Khan, S. U. D. Synthesis of Mater. 2008, 155, 39−44.

9598 https://doi.org/10.1021/acsomega.1c00143
ACS Omega 2021, 6, 9587−9599
ACS Omega http://pubs.acs.org/journal/acsodf Article

(66) Qu, J.; Zhang, Q.; Xia, Y.; Cong, Q.; Luo, C. Synthesis of
carbon nanospheres using fallen willow leaves and adsorption of
Rhodamine B and heavy metals by them. Environ. Sci. Pollut. Res.
2015, 22, 1408−1419.
(67) Wang, S.; Li, H.; Xu, L. Application of zeolite MCM-22 for
basic dye removal from wastewater. J. Colloid Interface Sci. 2006, 295,
71−78.
(68) Vijayakumar, G.; Tamilarasan, R.; Dharmendirakuma, M.
Adsorption, Kinetic, Equilibrium and Thermodynamic studies on the
removal of basic dye Rhodamine-B from aqueous solution by the use
of natural adsorbent perlite. J. Mater. Environ. Sci. 2012, 3, 157−170.
(69) Motahari, F.; Mozdianfard, M. R.; Salavati-Niasari, M.
Synthesis and adsorption studies of NiO nanoparticles in the presence
of H(2)acacen ligand, for removing Rhodamine B in wastewater
treatment. Process Saf. Environ. Prot. 2015, 93, 282−292.
(70) Marcano, D. C.; Kosynkin, D. V.; Berlin, J. M.; Sinitskii, A.;
Sun, Z.; Slesarev, A.; Alemany, L. B.; Lu, W.; Tour, J. M. Improved
synthesis of graphene oxide. ACS Nano 2010, 4, 4806−4814.
(71) Kumar, G.; Mogha, N. K.; Kumar, M.; Subodh; Masram, D. T.
NiO nanocomposites/rGO as a heterogeneous catalyst for imidazole
scaffolds with applications in inhibiting the DNA binding activity.
Dalton Trans. 2020, 49, 1963−1974.
(72) Subodh; Mogha, N. K.; Chaudhary, K.; Kumar, G.; Masram, D.
T. Fur-imine-functionalized graphene oxide-immobilized copper oxide
nanoparticle catalyst for the synthesis of xanthene derivatives. ACS
Omega 2018, 3, 16377−16385.
(73) Kumar, G.; Kant, A.; Kumar, M.; Masram, D. T. Synthesis,
characterizations and kinetic study of metal organic framework
nanocomposite excipient used as extended-release delivery vehicle for
an antibiotic drug. Inorg. Chim. Acta 2019, 496, 119036.
(74) Ming, Y.; Purewal, J.; Liu, D.; Sudik, A.; Xu, C.; Yang, J.;
Veenstra, M.; Rhodes, K.; Soltis, R.; Warner, J.; Gaab, M.; Müller, U.;
Siegel, D. J. Thermophysical properties of MOF-5 powders.
Microporous Mesoporous Mater. 2014, 185, 235−244.

9599 https://doi.org/10.1021/acsomega.1c00143
ACS Omega 2021, 6, 9587−9599

You might also like