You are on page 1of 14

Journal of Materials Science: Materials in Electronics

https://doi.org/10.1007/s10854-019-02324-7

Covalently linked porphyrin‑graphene oxide nanocomposite:


synthesis, characterization and catalytic activity
Altaf Ahmed1 · Gita Devi1 · Ashu Kapahi1 · Sujata Kundan2 · Sapna Katoch1 · Gauri D. Bajju1 

Received: 7 July 2019 / Accepted: 9 October 2019


© Springer Science+Business Media, LLC, part of Springer Nature 2019

Abstract
The work presented in this paper illustrates a synthetic endeavour devoted to the design and modification of an important
donor–acceptor porphyrin-graphene oxide (Por-GO) nanocomposite. The resulting novel Por-GO architecture was thoroughly
characterized with respect to their spectroscopic and catalytic properties. The functionalization of GO is mediated by an
axial ligand on the respective metalloporphyrin having an amide linker between the graphene oxide and the porphyrin moi-
ety. The successful functionalization of GO with axial ligand modified Ni(II)-porphyrin forming covalently linked Por-GO
nanocomposite with good dispersion stability in organic solvents was confirmed by various spectroscopic techniques viz.,
Fourier Transform Infrared Spectroscopy, Ultraviolet Visible Absorption Spectroscopy, Powder X-ray Diffraction, Scanning
Electron Microscopy and High-resolution Transmission Electron Microscopy. Finally, the catalytic reduction of 4-Nitrophenol
to 4-Aminophenol by sodium borohydride (­ NaBH4−) as reductant and synthesized nanocomposite as a catalyst was monitored
spectrophotometrically. The aqueous solution of 4-Nitrophenol is yellow in color and shows intense absorption in the visible
range. The completion of reduction reaction was confirmed by the change of yellow color to completely colourless solution
corresponding to the reduced product (4-Aminophenol). Furthermore, the complete disappearance of the absorption peak
related to 4-Nitrophenolate and emergence of a new peak exhibiting large hypsochromic shift corroborate the formation of
4-Aminophenol. The whole process of reduction completes within an hour as evident by time-dependent UV–Vis spectra.
The Por-GO nanocomposite exhibit excellent catalytic activity.

1 Introduction hydrophilic and water-soluble [6], and provides a handle for


the chemical modification of graphene using known carbon
Graphene, an allotrope of carbon, is one of the thinnest and surface chemistry [7]. Moreover, graphene oxide (GO) has
strongest materials in the universe and considered as the large π-electron network and edges provide additional cata-
most stunning nanomaterial owing to its fabulous physical lytic sites ideal for electrocatalytic applications. However,
and chemical properties [1]. It consists of a two dimensional GO is electrically insulating because of the discontinuous
one atom thick sheet of graphite having a honeycomb lattice ­sp2 bonding network and is thus unfavorable for the building
of ­sp2-carbon atoms [2]. Graphene is an excellent electron of optoelectronic devices [5, 8]. The electrical conductivity
acceptor [3] which can be easily functionalized with other and electron transfer properties of graphene can be partly
molecules and therefore, presents a set of properties like reinstated on moving from GO to reduced graphene oxide
mechanical rigidity, optical transparency, high chemical and (RGO). RGO has comparatively more electrochemical activ-
thermal stability and electrical transport [4]. The presence ity than GO because of the fewer oxygen functionalities on
of oxygen-containing groups like epoxides, hydroxyl and the edges. Implanting organic framework onto the surface
carboxylic acid [5] in graphene oxide renders it strongly of RGO proceed with retention of the structural integrity
of the RGO framework, thereby purging the problematical
* Gauri D. Bajju loss of electronic structure [9]. The unique two dimensional
gauribajju@gmail.com structures with its high specific surface area have made gra-
phene an attractive photocatalyst as well as ideal support
1
Department of Chemistry, University of Jammu, Jammu, for constructing new type of graphene-based photocatalysts
Jammu and Kashmir 180006, India
for photocatalytic reactions [10]. Graphene sheets can be
2
Department of Chemistry and Chemical Sciences, Central decorated with different functional groups or nanomaterial
University of Jammu, Jammu and Kashmir 180011, India

13
Vol.:(0123456789)
Journal of Materials Science: Materials in Electronics

to form graphene composites. The incorporation of gra- Herein, we report an approach for covalent grafting of a
phene into the composite can offer versatile characteristics metalloporphyrin over graphene oxide surface. The amino
providing a new opportunity for designing and developing group (–NH2) present on the axial salicylate ligand of the
next-generation catalyst. Chemical functionalization of RGO porphyrin was used to covalently link onto RGO surface by
may therefore potentially pave the way for their utilization the amide reaction between the amino group of porphyrin
in various realistic applications. complex and carboxyl (–COOH) group of RGO. The result-
Porphyrins on the other hand are regarded as one of ing nanocomposite was well characterized by various spec-
the most investigated group of macrocycles owing to their troscopic techniques. The catalytic activity of the resulting
exceptional photochemical and photophysical properties [11, porphyrin-graphene oxide nanocomposite was also evalu-
12]. Porphyrin consists of an 18 π-conjugated electronic ated spectrophotometrically for the catalytic reduction of
system moving in a square-planar circuit, giving rise to the 4-Nitrophenol. Furthermore, we discuss here the possible
aromatic character of the ring and consequently for its stabi- mechanism of the reduction reaction.
lization. Porphyrins are particularly amenable to the design
of complex and robust architectures because of their rigid
framework. In addition, porphyrin has an affinity to accom- 2 Experimental sections
modate with almost all the metals of the periodic table to
form various metalloporphyrins and subsequently used for 2.1 Materials
numerous purposes ranging from solar energy harvesting
to photodynamic therapy (PDT), etc. [13]. The extensive All reagents used in the experiment were of analytical
18 π-electron conjugated porphyrins have been widely stud- grade. Pyrrole was used as received and chloroform was
ied for the functionalization of graphene oxide on the basis first dried over C
­ aH2 and then distilled. Nickel(II) acetate
of π–π interactions between them. Linking porphyrins to [Ni(OAc)2·4H2O] was procured from Sigma Aldrich, para-
material systems can improve the photochemical and redox chlorobenzaldehyde ­(C7H5OCl) and ortho-nitrobenzalde-
properties of the systems [14]. Axial coordination of metal- hyde ­(C7H5NO3) were purchased from Himedia Labs and
loporphyrins however, offers a more versatile approach to used as received. 5-Aminosalicylic acid was procured from
the rapid preparation and photophysical characterization of SISCO Research Laboratories Pvt. Ltd. Mumbai-India.
self-assembled porphyrin-graphene oxide nanocomposites Silica gel (60–120 mesh) procured from Merck, Germany,
[15]. Anchoring graphene with porphyrin through covalent was used for column chromatography. Graphite flakes and
or non-covalent bonds can result in effective heterogeneous potassium permanganate (­ KMnO4) were purchased from
interfacial electron transfer and long-distance transport at the Alfa Aesar and were used as received without further puri-
surface junction and therefore, it is expected that the cata- fication. Ethanol, sulphuric acid, phosphoric acid, hydrogen
lytic activity of the resultant nanohybrid can be tailored. The peroxide, hydrazine hydrate, thionyl chloride ­(SOCl2) and
coordination chemistry of axially ligated metalloporphyrins triethylamine ­(Et3N) were all supplied by Himedia Chemical
with polydentate salicylic acid derivatives has provided a Reagent Company.
vast field for the building of supramolecular entities [16].
There is a paucity of reported work consisting of covalently 2.2 Synthesis
linked Ni(II)-porphyrin to the graphene oxide-mediated by
axial ligands, although reports of various metalloporphyrins The synthesis process involves the following three steps.
linked through peripheral groups can be found in the litera-
ture [14]. Ni(II)-porphyrins have received attention as recep- 2.2.1 Synthesis of graphene oxide and its derivatives
tors for various oxygen donors and their subsequent linking
to carbonaceous materials. Porphyrin-based graphene oxide 2.2.1.1  Preparation of  graphene oxide (GO) from  graph‑
(GO) nanocomposites are fascinating since they merge the ite flakes Graphite flakes were used for the synthesis of
exceptional properties of two unique materials. Porphyrin graphene oxide (GO) following Hummers and Offeman’s
itself exhibit excellent catalytic chemistry [17]. Combination method with modification [21, 22]. The technique involves
of the unparalleled properties of graphene and the porphy- the mixing of concentrated ­H2SO4/H3PO4 in 9:1 (81:9 mL)
rin on its surface led to improvement the composite perfor- ratio to a mixture of 1.25 g of graphite flakes and 9.0 g of
mance. In fact the use of graphene as a support for porphyrin ­KMnO4, producing a small exotherm (35–40  °C). Subse-
led to a measureable enhancement of the electronic proper- quently, the reaction mixture was stirred for 12 h at 50 °C,
ties and synergistic interactions between the porphyrin and cooled to room temperature and poured into ice-cold water
the graphene. Many reports of graphene oxide functional- (100  mL) containing 30% H ­ 2O2 (1.25  mL) and finally fil-
ized with porphyrins and their metal complexes are available tered. The filtrate was centrifuged (3000 rpm for 2 h) and the
in literature [7, 18–20]. supernatant was decanted away. The residual solid material

13
Journal of Materials Science: Materials in Electronics

was then washed twice in succession with 50 mL of water, 2.2.2 Synthesis of axially ligated metalloporphyrin
50  mL of dilute 30% HCl, and 50  mL of ethanol. Finally ­[NH2SA‑NiPor]
washed with diethyl ether and the solid so obtained were
dried in vacuum overnight at room temperature yielding the A brief outline of the procedure is summarized in Scheme 1.
desired product. The procedure consists of first preparing metal-free porphy-
rin precursor followed by metalation and subsequent axial
ligation in three successive steps as below;
2.2.1.2  Preparation of  reduced graphene oxide (RGO)
from graphene oxide  In a 250 mL round-bottom flask con- 2.2.2.1  Step‑(i) Synthesis of  5,10,15‑Tris(4‑Chlorophenyl)
taining 100 mL of distilled water, graphene oxide (1.0 mg) ‑20‑(2‑Nitrophenyl)porphyrin (Por) The synthesis and
was added and kept on a magnetic stirrer. After sonication characterization of ­A3B-type unsymmetrical free-base por-
for 30 min, 10 mL of 50% hydrazine-hydrate was gradually phyrin were carried out according to our previously pub-
added to the resulting reaction mixture and refluxed for 4 h lished procedure [23].
at 100  °C. The solution mixture so obtained was filtered, Characterization Anal. calcd. for C44H26N5Cl3O2 (%): C
washed thoroughly with anhydrous THF and dried under 69.26, H 3.43, N 13.94. Found: C 69.59, H 3.70, N 14.12,
vacuum for 24 h. UV–Vis ­(CHCl3): 460, 580, 623, 657 and 729 nm. 1H NMR
(400 MHz, ­CDCl3): δ (ppm) = − 2.70 (2H, s, NH), 7.29
(2H, d, Ar-NO2H4,5), 7.75(6H, d, A ­ rH3,5), 8.11 (2H, d, Ar-
2.2.1.3  Acylation of  reduced graphene oxide; (GO‑COCl)  NO2H3,6), 8.16 (6H, d, ­ArH2,6), 8.80 (2H, d, βHa), 8.87 (4H,
Thionyl chloride ­(SOCl2) was added in reduced graphene s, βHc), 8.91 (2H, d, βHb). ESI–MS ­(CH3OH): m/z calcd. for
oxide (1.0 g) under an inert ­(N2) atmosphere in a fuming C44H26N5Cl3O2: 763.07; found 764.01 ([M + H]+), Rf = 0.71.
hood. After 5 min, 10 mL of DMF was added into the result-
ing mixture and refluxed at 70 °C for 24 h. The solution was 2.2.2.2  Step‑(ii) Synthesis of  5,10,15‑Tris(4‑chlorophenyl)
then filtered and washed thoroughly with anhydrous THF ‑20‑(2‑nitrophenyl)nickel(II)‑porphyrin (NiPor) The syn-
and dried under vacuum for 24 h. At the end of the reaction, thesis of the corresponding metalloderivative of free-base
excess ­SOCl2 and solvent were removed by distillation. porphyrin was carried out by treating Por with the Ni(II)-
acetate dissolved in methanol and chloroform [24].

Scheme 1  Synthesis of Axially ligated Ni(II)-porphyrin

13
Journal of Materials Science: Materials in Electronics

Characterization Anal. calcd. for C 44H 24N 5Cl 3NiO 2 2.3 Catalytic activity determination
(%): C 64.47, H 2.95, N 12.97. Found: C 64.19, H 3.10, N
13.23, UV–Vis ­(CHCl3): 445, 618 and 675 nm. 1H NMR The catalytic activity testing of porphyrin-graphene oxide
(400 MHz, ­CDCl3): δ (ppm) = 9.37 (d, βHa), 8.97 (s, βHc), (Por-GO) nanocomposite was carried out by first preparing
9.01 (d, βH b), 7. 73 (6H, d, Ar-H 3,5), 7.28 (2H, d, Ar- an aqueous solution of both the reactants; 4-Nitrophenol
NO2H4,5), 8.15 (6H, d, Ar-H2,6), 8.09 (2H, d, Ar-NO2H3,6). and sodium borohydride (­ NaBH4) in the molar concentra-
ESI–MS ­( CH 3 OH): m/z calcd. for C 44 H 24 N 5 Cl 3 NiO 2 : tion of 1 mM and 0.3 M respectively at room temperature
819.75; found 821.03 ([M + H]+), Rf = 0.52. and neutral pH conditions. Two separately prepared fresh
solutions of both the reactants (each 1 mL) were mixed by
2.2.2.3  Step‑(iii) Synthesis of  5″‑Aminosalicylato‑5,10,1 magnetic stirring and transferred to a quartz cuvette/cell.
5‑tris(4‑Chlorophenyl)‑20‑(2‑Nitrophenyl)nickel(II)‑por‑ In one of the solution, 5 mg of prepared catalyst (Por-GO)
phyrin (NH2SA‑NiPor)  The synthesis of axial salicylate was loaded into the resulting reaction mixture placed in the
ligated NiPor was carried out according to our previous cuvette to start the reaction and immediately placed in the
work [25, 26]. cell holder of the spectrophotometer. From the second blank
Characterization Anal. calc. for C 51 H 30 N 6 Cl 3 NiO 5 solution of 4-Nitrophenol and N ­ aBH4, 5 mL of the solu-
(%): C 63.03, H 3.11, N 8.65. Found: C 62.89, H 2.97, tion was taken and the absorption spectrum was recorded
N 8.27, UV–Vis (­CHCl 3 ): 450, 570 and 673  nm. 1 H on UV–Vis spectrophotometer which is taken as read-
NMR (400 MHz, C ­ DCl3): δ (ppm) = 9.55 (d, βHa), 9.09 ing at 0 min. After every 4 min absorption spectra were
(s, βH c), 9.17 (d, βH b) 8.01 (6H,d, Ar-H 3,5), 7.75 (2H, recorded by taking 5 mL of reactant solution containing Por-
d, Ar-NO 2H 4,5), 8.78 (6H, d, Ar-H 2,6), 8.43 (2H, d, Ar- GO catalyst. The same procedure was repeated until one
NO2H3,6), 7.07 (d, H ­ 4″), 7.39 (d,H3″), 7.41 (s,H6″). ESI–MS hour, in order to measure the residual concentration of the
­(CH3OH): m/z calc. for C51H30N6Cl3NiO5: 971.88; found 4-Nitrophenol. The absorbance at λ = 400 nm in the UV–Vis
971.03 ([M + H]+), Rf = 0.67. absorption spectra was used to monitor the progress of the
reduction reaction i.e. the conversion of 4-Nitrophenol into
4-Aminophenol by recording the time-dependent absorption
2.2.3 Synthesis of covalently linked porphyrin‑graphene spectra at ambient temperature.
oxide nanocomposites (Por‑GO)
2.4 Characterization
Carboxylic acid groups situated at the periphery of
GO were utilized to covalently graft porphyrin through Thin layer chromatography was performed on pre-coated
amide linkages [27]. This is achieved by the function- TLC Silica ­gel 60 on aluminium sheet, purchased from
alization of acylated graphene oxide with axially ligated Merck. UV–Vis spectra were recorded on a T90 + UV–Vis
Ni(II)-porphyrin. In this method, GO-COCl (60 mg) and spectrophotometer in the range 350–700 nm. Column chro-
­NH2SA-NiPor (60 mg) was taken in a 250 mL round bot- matography was carried out on silica gel (100–200 mesh)
tom flask. The reaction mixture was dissolved in 30 mL purchased from Merck. The FTIR spectra were recorded
of N, N-dimethylformamide (DMF) in the presence of tri- on a Fourier Transform Infrared (FTIR Shimadzu Pres-
ethylamine ­(Et3N, 3 mL) and was allowed to react under tige-21) spectrophotometer. X-ray diffraction patterns of
reflux at 100 °C for 72 h in an inert atmosphere. Conse- prepared samples were recorded by using an XPERT-PRO
quently, the mixture was sonicated, producing a uniform diffractometer system with monochromatic Cu K ­ α radiation
black dispersion. The reaction mixture was then cooled to (λ = 1.54060 Å), at a scanning rate of 12 °C/min in the 2θ
room temperature and poured into 300 mL of diethyl ether range of 10° to 70°. The SEM imaging of the prepared sam-
to precipitate the product. The precipitates were separated ples was carried out by using Nova Nano SEM-450 (JFEI
by centrifugation (5000 rpm) and the supernatant contain- Company of USA). The TEM imaging of the synthesized
ing unreacted porphyrin as impurities were discarded. The complexes was carried out by using Tecnai G2 20 S-TWIN
precipitates were rinsed thoroughly over and again in five (FEI Company of USA).
washing successions. Each time sonication, filtration and
re-suspension of the precipitates in THF (60  mL) was
done. Finally, the product was washed several times with 3 Results and discussion
­CHCl3 and water to remove the excess ­Et3N·HCl and dried
under vacuum. TLC was used to ensure the absence of any The coordination of 5″-Aminosalicylic acid to Ni(II)-por-
unreacted porphyrin in the final product. The procedure phyrin occurs through the carboxylic oxygen via depro-
adopted for the synthesis of covalently-linked GO-porphy- tonation leaving behind the aromatic –OH and an amino
rin composites is outlined in Scheme 2. group (–NH2) at 2″ and 5″-Carbon respectively, free to

13
Journal of Materials Science: Materials in Electronics

Scheme 2  Schematic representation for the synthesis of covalently-linked axially ligated porphyrin-GO (Por-GO) nanocomposite

interact with other ligands. Since the aromatic –OH is steri- groups [28]. Majority of the oxygen-containing groups on
cally hindered and –NH2 spatially more oriented for further the surface of GO will disappear after reduction. The most
coordination, therefore, it is expected that a combination of prominent absorption band displayed by GO is a broad and
oxyphilic Ni(II)-porphyrin would be suitable for construc- intense band at 3178 cm−1 that can be assigned to O–H
tion of supramolecular architectures with graphene oxide stretching vibration. Another characteristic band in the FTIR
via amide-linkages. Axial ligation of the metalloporphyrins, spectra of GO at 1730 cm−1 is attributed to the stretching
therefore, affords an additional means for assembly or organ- vibrations of C = O of carbonyl and carboxyl groups [29].
ization. Thus, the linking of 5″-aminosalicylato-5,10,15- Band appearing at 1624 cm−1 is assigned to the bending
tris(4-chlorophenyl)-20-(2-nitrophenyl)nickel(II)-porphyrin vibration of OH-group whereas, ν(epoxy) band appears at
­(NH2SA-NiPor) (electron donor) to the reduced graphene 1170  cm−1. A band attributed to stretching vibration of
oxide (electron acceptor) via an amide linkage results in C–O–C (epoxy group) appears at 1054 cm−1. The bands at
a perfect donor–acceptor nanohybrid. The attachment of 1407–1330 cm−1 are the C–O absorption of carboxyl. From
organic molecules to graphene oxide has made the Por-GO the FTIR analysis, it can be assumed that the presence of
soluble in DMF and other polar solvents. The characteri- carboxyl, hydroxyl, epoxy and carbonyl groups in GO lead
zation of the synthesized Por-GO nanocomposite has been to an increase of interlayer distance.
discussed in subsequent sections. FTIR measurement results of different samples are shown
in Fig. 2 for the rationale of characterizing the difference
3.1 FTIR spectroscopy between reduced graphene oxide and Por-GO nanocompos-
ites. It can be seen that the RGO still contains some oxygen
FTIR spectrum was used to investigate the structure and functionalities as evident from the peak around 1720 cm−1.
nature of functional groups on the surface of GO. The FITR The less intense peak corresponding to carbonyl or car-
spectrum shown in Fig. 1 reveals that GO has many oxygen- boxyl group indicates that only fewer oxygen functionali-
containing groups such as hydroxyl, epoxide and carboxyl ties still remains on the surface of graphene oxide even after

13
Journal of Materials Science: Materials in Electronics

Fig. 1  FTIR spectra of graphene


oxide (GO)

attributed to the stretching vibrations of C–H of GO, RGO


and Por-GO nanocomposite. The NiPor and N ­ H2SA-NiPor
spectrum shows many characteristic bands. The presence of
a band around 2900 cm−1 corresponding to the stretching
vibrations of C–H bonds of the phenyl. Moreover, the band
at 1550 cm−1 is attributed to C–C stretching. A broad band
at 3400 cm − 1 is attributed to the N–H bond stretching of
the axial 5-Aminosalicylate ligand [30].
The covalent functionalization of GO with porphyrins
results into the appearance of a new peak ~ 1600 ­cm−1
related to the C = C vibrations of porphyrins along with the
shift of peak at 1730–1640 c­ m−1 for Por-GO nanocompos-
ites corresponding to the C = O stretching vibration of the
amide group. The peak of the amide C–N stretching vibra-
tion appears at 1250 c­ m−1. The peak present at ~ 1707 cm−1
is assigned to the bending vibration of the C = N of the por-
Fig. 2  FTIR spectra of RGO, NiPor, ­
NH2-SA-NiPor and Por-GO
phyrin ring. These results clearly point out that in graphene
nanocomposite oxide–porphyrin composites, the porphyrin molecules are
covalently linked to the graphene oxide via amide-linkages.
Moreover, all other frequencies corresponding to GO-free
reduction. The absorption intensity of the O–H band of GO porphyrin complex i.e. NH-2SA-NiPor are also presented by
is stronger than corresponding RGO and Por-GO nanocom- the nanocomposite as well.
posites, it is proposed that the graphene oxide contains a
greater number of O–H groups on the surfaces due to oxida- 3.2 UV–Vis absorption spectroscopy
tion. Although the graphene oxide gets reduced during the
process of reduction, the FTIR spectrum of RGO even shows The electronic absorption spectra of the free base porphy-
some characteristic bands of GO such as stretching vibra- rin (Por) exhibit two characteristic types of bands [31]; an
tions due to OH and C = O. The main characteristic bands of exceptionally intense Soret (B-band) band molar absorp-
RGO are located at 1339, 1581, and 1632 cm−1, which are tivities, ε, on the order of ­105 M−1 cm−1, can be found in
attributed to stretching/deformation of C–OH and aromatic the near-UV or shorter-wavelength regions (400–450 nm)
C = C stretching, respectively. It has been reported that the and four weak bands, the Q-bands with one order of mag-
FT-IR spectra of reduced graphene oxide is very similar to nitude lower molar absorbance, assigned to the π–π* transi-
that of graphite. Some peaks observed around 2900 cm−1 are tion between bonding and anti-bonding molecular orbitals

13
Journal of Materials Science: Materials in Electronics

­(S0 → S1), appear in the visible region of the spectrum or monotonically from the UV to the visible region and a small
longer-wavelength range (500–700 nm). However, the shoulder band around 315 nm. These bands are attributed
UV–Vis spectra of NiPor exhibited blue-shifts in B-band to the π–π* transitions of aromatic C = C band and n–π*
which is attributed to the metal d­ π ­(dxz and ­dyz) to porphy- transitions of C = O band [22], respectively. However, the
rin π* (M → L) back-bonding and the four Q-bands of free- reduced graphene oxide (RGO) exhibit a red-shifted UV–Vis
base porphyrin collapse to two bands. This dramatic effect is absorption spectrum (λmax = 276 nm) compared to GO, the
ascribed to the enhancing of the D ­ 2h symmetry of the free- shoulder band at 315 nm disappears and the absorbance in
base porphyrin to D­ 4h on metalation (Fig. 3a). Moreover, the the whole spectral region decreases. The bathochromic shift
Por display ‘etio-type’ Q-band spectrum in which the rela- indicates that the GO was efficiently reduced. Certainly, the
tive intensities of these bands are such that IV > III > II > I reduction in oxygen functionalities led to slight diminishing
as shown in Fig. 3a (Por x 10). The extremely low intensity in the absorption intensity, thus the electrons can easily be
of Q–I band may be due to the unsymmetrical nature of the excited at lower energy and a slight shift is noted after the
porphyrin causing distortion. reduction of GO sheets.
The optical absorption spectra recorded in DMF for Conversely, the absorption spectrum of free porphy-
graphene oxide (GO), reduced graphene oxide (RGO) rin complex ­(NH 2SA-NiPor) in which the delocalized
before and after its covalent interactions with porphyrin π-electrons shows very strong π–π* absorption and exhibit
molecule are shown in Fig. 3b. Graphene oxide displays a strong soret absorption at 419 nm and two weak Q-bands
a broad absorption signal [32, 33] at 265 nm decreasing at 518 and 549 nm. The nanocomposite (Por-GO) exhibits
a broad absorption at 283 nm, which is the corresponding
RGO peak with a red-shift of 7 nm. Moreover, in the Por-GO
(a)
nanocomposites, a similar band is also observed at 425 nm,
which corresponds to the soret band of the porphyrin moiety
and obviously a red-shift of 6 nm is observed which is attrib-
uted to the covalently grafted porphyrin moiety. However,
the Q-bands as displayed by the free porphyrin complex at
518 and 549 nm get shortened, red-shifted and flattened to
the baseline in the corresponding nanomaterial [34] and only
a broader hump is observed at 550 nm.
The red-shifts in the soret band [35] from 419 to 425 nm
for the nanocomposites can also be ascribed to the adsorp-
tion of porphyrin complex on GO induced by π–π stack-
ing interactions between the porphyrin moiety and the GO.
These results corroborate not only the linkage of porphyrin
with the GO sheets but also indicate that attachment of the
porphyrin moiety has perturbed both the ground state elec-
tronic state of the graphene oxide as well as of the porphyrin
(b)
part.

3.3 Powder X‑ray diffraction

The structural characterization of the nanoparticles has been


carried out by Powder X-ray diffraction technique using Cu
Kα radiation [36]. The powder X-ray diffraction (PXRD)
pattern of graphene oxide is shown in Fig. 4. A sharp peak
(2θ) at 10.15° (001) and a weak peak at 42.05° (002) cor-
responding to layer-to-layer distance (d–spacing) spacing of
8.27 and 3.47 Å respectively, are characteristic diffraction
peaks for graphene oxide. The diffraction peak at 2θ = 10.15°
is indistinct, which suggests that the graphite is exfoliated to
graphene oxide and reduced graphene oxide; the sectional
oxygen-containing groups were removed during reduction
Fig. 3  a UV–Vis absorption spectra of Por, NiPor and Q-Bands of
Por (x 10) in ­CHCl3. b UV–Vis absorption spectra of GO, RGO, treatment. Some weak and broad peak in the range of (2θ)
­NH2SA-NiPor and Por-GO nanocomposites 20° corresponds to (100) with the d-spacing 3.17 Å indicates

13
Journal of Materials Science: Materials in Electronics

broader indicating typically amorphous nature. Moreover, in


the XRD pattern of Por-GO nanocomposites, the presence of
smaller peaks around 2θ = 13o indicates oxygen-containing
groups still exist in the nanocomposites through the oxygen-
containing groups were partly removed during exfoliation.
The presence of crystalline domains could enhance the
charge transport and stabilize the amorphous Por-GO dur-
ing charging and discharging. The PXRD profile confirmed
the amorphous character of cross-linked porphyrin-graphene
oxide nanocomposites. Therefore, the cross-linked Por-GO
is an amorphous and nanoporous polymer bearing covalently
linked porphyrin functionalities in the nanocomposites.
These results corroborate the formation of a new product
as a result of covalent linkage of porphyrin moiety with the
GO surfaces.
Fig. 4  PXRD pattern of graphene oxide (GO)

3.4 Scanning electron microscopy

Morphology and topography of graphene oxide have been


investigated by scanning electron microscopy (SEM). The
graphite powder is composed of a number of flake-like
layers. However, after the oxidation process, these graph-
ite layers were exfoliated and subsequently intercalated by
epoxy, carbonyl, carboxyl and hydroxyl functional groups of
graphene oxide. The SEM images of synthesized graphene
oxide and its covalently linked Por-GO nanocomposites are
shown in Fig. 6. The SEM micrograph shows that the gra-
phene oxide has a layered structure [41] with a large number
of cavities and free space, which affords ultrathin and homo-
geneous films. It is observed from the images that crumpled
and rippled structures are there which are due to resulting of
deformation upon the exfoliation and restacking processes.
Figure 6 shows that the graphene sheets are entangled with
each other and also the single or few-layer graphene oxide
Fig. 5  PXRD pattern of Por-GO nanocomposite
sheets have lots of wrinkles.
The scanning electron micrographs for the Por-GO nano-
the presence of some under oxidized graphitic material. The composites reveal the expected sheet morphology over a few
results obtained are in accordance with the literature [37]. micron length scales. The graphene sheets in the nanocom-
These peaks vary, depending on the preparation method and posites appear somewhat more wrinkled compared to the
the number of layers in the graphite gallery. neat the GO sheets. A stacked lamellar morphology with
However, the parent graphite exhibits a peak at 26.31° graphene sheets was also observed in the Por-GO nanohy-
(2θ) [38]. This downshift of XRD-signal from 26.31° (2θ) brid. On account of functionalization, GO exhibits a three-
to 10.5° (2θ) after oxidation is an indication of decreasing dimensional network of randomly oriented sheet-like struc-
graphene layer number in graphene oxide (GO) due to exfo- tures with a wrinkled texture and hierarchical pores with
liation with respect to parent graphite, suggesting the pres- wide size distribution. The SEM-micrograph for the Por-
ence of few-layered graphene sheets [39]. GO nanohybrid materials also demonstrates a homogeneous
For the Por-GO Nanocomposites (Fig. 5), a comparatively system. It is clearly seen from the top-view image that the
broader peak is observed at 12.1° (2θ) corresponding to a composites have a uniform and dense surface, composed of
d-spacing of 7.52 Å. This indicates that the interlayer spac- plate-like particles with slightly scrolled edges ranging from
ing was maintained due to the intercalation of porphyrin several tens of nanometers to several hundreds of nanom-
and the successful formation of the Por-GO nanocomposites eters in the lateral dimension. Some GO flakes of composites
[40]. The peaks of the nanocomposites are comparatively fold together and represent crinkly sheets. The surface is

13
Journal of Materials Science: Materials in Electronics

Fig. 6  a–b SEM images of graphene oxide c ­NH2SA-NiPor and d Por-GO nanocomposite

much rougher than that of GO which can be attributed to the ­NH2SA-NiPor and GO. The images reveal a uniform distri-
covalent linkage of porphyrin to the GO sheet [42]. bution of porphyrin nanocomposites on the surface of gra-
phene sheets making rough (granular) surfaces.
3.5 Transmission electron microscopy

The morphological and structural features and extent of 4 Catalytic activity of Por‑GO
covalent linking of Por-GO nanocomposite as investigated nanocomposite
by SEM analysis were further supported by TEM micro-
graphs. Figure 7 shows the selected high-resolution TEM Nitroaromatic compounds and their derivatives are key
(HR-TEM) micrographs for GO and free porphyrin with organic pollutants created as by-products or intermediates
their corresponding nanocomposites. The HR-TEM image in a number of industrial processes [44, 45]. Plentiful of
of GO has few micrometer lengths with fold however, that of nitroaromatic compounds have been discharged into the riv-
RGO reveals a wrinkled, curled and corrugated morphology ers and soil amid the over use of dyes, explosives and pes-
typical of few-layered structure [43]. It has been found that ticides in industry, causing severe water and soil pollution
the graphene flakes have wrinkled surfaces. Furthermore, in [46]. Regrettably, among nitro-compounds, 4-Nitrophenol
the TEM image GO shows the layer-by-layer stacked struc- is the most typical toxin and insurmountable organic pollut-
ture and have a wrinkled paper like morphology. Such mor- ant, distressing the eco-systems of both humans as well as
phological changes can be attributed to the increased forma- animals and hence promoting a variety of ailments. 4-Nitro-
tion of phenolic and epoxy functional groups on the basal phenol with other derivatives is a familiar by-product of the
plane of GO. The curled and overlapped nanosheets can be chemical industry. The manufacture of pesticides, herbi-
clearly observed. The existence of small nanoparticles (dark cides, pharmaceuticals, synthetic dyes, cosmetics, photo-
spots) [5] distributed on the wrinkled graphene sheets of graphic chemicals, etc. produces 4-Nitrophenol as the side-
Por-GO Fig. 7d confirms the presence of N ­ H2SA-NiPor, and product. 4-Nitrophenol notably is well-known to cause lethal
further supports the presence of covalent bonding between effects in the biological systems particularly it affects the

13
Journal of Materials Science: Materials in Electronics

Fig. 7  HR-TEM micrographs of a GO b RGO c ­NH2SA-NiPor and d Por-GO nanocomposites

liver, kidney and blood of both animal and human and dam- Use of semiconductor nanomaterials such as ZnO, ­Cu2O and
age to the central nervous system owing to their severe tox- ­TiO2 and various methods (e.g. photocatalysis and chemical
icity in water and soil. Therefore, its degradation into less/ catalysis) have been reported and even succeeded in its deg-
non-toxic small molecules and in turn thwarting mischief radation or reduction [54]. But no efficient method has been
is a crucial task and has become an active area of research developed so far to deal with the problem of slow reduction
in recent years. Likewise, the reduction of 4-Nitrophenol is of 4-Nitrophenol. The difficulty to degrade 4- Nitrophenol is
particularly indispensable in pharmaceutical industries for due to the high stability and low solubility of the substrate in
the manufacture of analgesic, antipyretic and primarily a water [55]. Most recently, porphyrin-based graphene oxide
key intermediate in the synthesis of paracetamol [47, 48]. nanocomposites have been intensely explored because of
Moreover, it also finds relevance as a photographic devel- their unique and tuneable electric, thermal, mechanical, opti-
oper, corrosion inhibitor, anticorrosion lubricant, drying cal properties and high catalytic activities.
agent [49, 50], etc.
In general, the catalytic conversion of 4-Nitrophenol to 4.1 Catalytic reduction of 4‑nitrophenol
4-Aminophenol is a crucial process, using excess sodium
borohydride ­( NaBH 4) as the reducing agent, not only To determine the catalytic behaviour of porphyrin-based
because 4-Aminophenol is less toxic than 4-Nitrophenol, graphene oxide nanocomposites, the catalytic reduction of
but also because 4-Aminophenol is in demand on the market 4- Nitrophenol was chosen as a model reaction. The aque-
in many industrial fields [51]. Disappointingly, the reduc- ous solution of 4-Nitrophenol is yellow in color that assists
tion process of 4-Nitrophenol to 4-Aminophenol is very to study the kinetics of the reaction spectrophotometrically.
sluggish by ­NaBH4 alone. Hence, the use of catalysts is The nitrophenolate ion at a higher alkaline pH (pH > 12.0)
necessary to enable electron transfer from donor ­BH4− to increases the spectrophotometric sensitivity of the method
acceptor 4-Nitrophenol [52]. Many practices have been in that can be gauged by a visible spectrophotometer only.
use to solve the pollution problems of 4-Nitrophenol (nitro- It has been reported that the original aqueous solution of
group removal or its conversion to non-toxic form) [53]. 4- Nitrophenol has a maximum absorption peak (λmax) at

13
Journal of Materials Science: Materials in Electronics

317 nm [56] and if no catalyst is added, the peak at 317 nm


in the UV–Vis spectra remains unchanged even after keep-
ing the reaction mixture for a long time. The uncatalyzed
reaction does not proceed favorably well in the forward
direction and hence Por-GO nanocomposite plays its role to
catalyze the reduction process. Upon addition of a freshly
prepared solution of ­NaBH4 solution, the absorption peak at
317 nm undergoes bathochromic shift (red-shift) to 400 nm
due to the formation of 4-Nitrophenolate ions, and mean-
while, the color of the mixture changes from light yellow to
dark yellow. The ­NaBH4 leads to an alkaline pH and the yel-
low solution of 4-Nitrophenolate remains stable for hours if
Fig. 8  UV − Vis spectra of the reduction process of 4-Nitrophenolate
no catalyst is added. The reduction reaction cannot proceed to 4-Aminophenolate catalyzed by Por-GO nanocomposite
in the forward direction without catalyst [57] even in the
presence of excess ­NaBH4 as evident from the almost con-
stant absorbance at 400 nm. However, it has been observed more reactant is left behind. The product, 4-Aminophenol
that upon addition of a catalytic amount (5 mg) of Por-GO devoid of any chromophore imparts no color to the aqueous
nanocomposite, the yellow color of the solution decolorized solution and hence does not hamper the spectrophotometric
within an hour. After a reaction time of about 50 min, no determination.
4-Nitrophenol could be detected by UV–Vis spectroscopy.
The reaction was carried out at room temperature. Schematic 4.2 Mechanism of catalytic reduction
representation of the procedure is described in Scheme 3.
The time-dependent UV–Vis absorption spectra of the It has now been established beyond suspicion that cataly-
catalytic reduction process of 4-Nitrophenol by Por-GO sis involving nitroarenes reduction depends not only on
nanocomposites in the presence of excess N ­ aBH4 at a cer- the size but also on the shape and surface of the catalyst
tain time interval is displayed in Fig. 8. As soon as the Por- particle as well. Therefore, the catalysts having greater sur-
GO nanocomposite (catalyst) is introduced into the reaction face area, porosity, facet selection and heterojunction etc.
mixture, the intensity of the main absorption peak around has been found to change the potential of a catalyst dra-
400 nm corresponding to organic pollutant starts diminish- matically. There exists a number of reported mechanisms
ing [58]. Simultaneously, a new peak around 300 nm starts for the reduction of niroarenes [59]. The reduction reaction
emerging and increases (indicated by arrows) in intensity of 4-Nitrophenol to 4-Aminophenol occurs in the presence
as a function of the reaction time. The augment of a new of a reductant, sodium borohydride (­ NaBH4) and a catalyst.
peak around 300 nm is attributed to the formation of 4-Ami- This reduction is ideally considered to proceed through a
nophenol, the subsequent products of the reduction reaction number of stages. Thus, it is corroborating a straight forward
of 4-Nitrophenol. The peak at 400 nm disappears completely process to produce only 4-Aminophenol without any side
within 50 min of reaction time. The yellow color of the solu- product. Thermodynamically the reduction of 4-Nitrophe-
tion converted to colorless, demonstrating that the 4-itro- nol to 4-Aminophenol (Eº = − 0.76 V) at ambient condition
phenol is completely changed into 4-Aminophenol and no is favourable as borohydride in water is a strong reductant

Scheme 3  Schematic represen-
tation of the catalytic reduction
of 4-Nitrophenol to 4-Ami-
nophenol by ­NaBH4

13
Journal of Materials Science: Materials in Electronics

(Eº = − 1.33 V). The reduction reaction in absence of Por- OH


GO catalyst does not proceed in the forward direction under
the experimental time scale because of the kinetic barrier. Por
This barrier is judiciously by-passed through the deployment 6H+
of catalyst and even catalysts with a huge support. In the Por
reduction process the electron relay takes place irreversibly NaBH4
[11]. Por NO2
e 6e
Exactly how Por-GO catalyst flank the surface atoms to
accommodate 4-Nitrophenol to 4-Aminophenol molecules OH

remain unclear. However, the mechanism of the reaction can


be understood in terms of Langmuir–Hinshelwood model
Por
[60]. The mechanism of catalytic reduction of 4-Nitrophe- Por
nol to 4-Aminophenol involves the diffusion and adsorption Por
of both the reactants [61]; 4-Nitrophenol and the reduct- NH2
ant ­NaBH4 to the surface of the Por-GO nanocomposite
[62]. The substrate anchors on the nanocatalyst sheet due
Fig. 9  Diagrammatic representation of the catalytic process of reduc-
to the presence of –OH and –NO2 groups. The substrate tion of 4-Nitrophenol to 4-Aminophenol facilitated by Por-GO nano-
gets uniformly distributed on the Por-GO hybrid catalyst composite
surface. The reaction medium is mostly basic as sodium
borohydride in aqueous medium splits to give borohydride
anion and sodium cation which makes the solution basic in enabling efficient contact between them. The schematic
nature. In the direct route of substrate reduction, the reduc- representation of the mechanism of catalytic reduction of
tion of 4-Nitrophenol to 4-Aminophenol occurs in a step- 4-Nitrophenol to 4-Aminophenol is shown in Scheme 4.
wise manner through the formation of 4-Nitrosophenol and After completion of the reduction reaction, the recyclabil-
4-hydroxylaminophenol as intermediates by using six elec- ity and reusability of the Por-GO catalyst was checked by
trons and six protons. The electron transfer occurs from the UV–Vis absorption spectroscopy. The catalyst was recov-
negatively charged B ­ H4− to 4-Nitrophenol [63]. It is note- ered by centrifugation followed by washing with metha-
worthy to mention here that the 4-hydroxylaminophenol is nol and dried at 50 °C. The recovered photocatalyst was
the first stable intermediate of the reduction reaction. Hence, reused without recharging for about four subsequent runs
the compounds that compete to undergo adsorption and des- under described experimental conditions. No significant
orption cycle with the fixed number of sites of the catalyst’s loss in activity was observed for the recycled catalyst, and
surface are 4-Nitrophenol, 4-Nitrosophenol, 4-hydroxylami- the reduction reaction time remained almost unchanged
nophenol and 4-Aminophenol. The adsorbed species then even after four recycling experiments, which confirmed the
reacts and finally the reaction product dissociate from the true heterogeneous nature of the developed photocatalyst.
surface of catalyst. The diagrammatic representation of the The absorption spectra of the Por-GO nanocomposite did
catalytic process is presented in Fig. 9. not show any change even after four recycling experiment.
The Por-GO support plays an active part in the cataly- These results corroborate not only that the covalent bonding
sis to give a synergistic effect. The RGO increases catalytic between the RGO and porphyrin complex as strong enough
activity of the Por-GO nanocomposite by facilitating elec- but also the porphyrin complex did not show any leaching
tron transfer to reactant molecule along with the extensive out from the surface of RGO during the reaction.
sites for the attachment. RGO, due to the presence of exten-
sive network of conjugated ­sp2 carbons provides excellent
mobility of electrons. Therefore, the synergistic catalytic 5 Conclusions
activity of the porphyrin and RGO can be explained in terms
of catalytic nature of Ni(II)-metal inside the porphyrin core A porphyrin-graphene oxide nanocomposite forming
whereas, the RGO provide greater surface area for anchor- donor–acceptor system has been reported. The resulting
ing the substrate to be reduced. RGO alone was also found supramolecular architecture was well characterized by
to catalyse the reduction reaction of 4-Nitrophenol but with various spectroscopic studies. The covalent functionali-
greater reaction time (≈ 5 h). Moreover, it is also projected zation of GO with porphyrin results into the appearance
that the π–π interactions [64] between the 4-Nitrophenol and of a new peak in the FTIR spectrum. The Soret band in
Por-GO nanocomposite helps the adsorption of 4-Nitrophe- the UV–Vis spectra of Por-GO nanocomposite, display a
nol on nanocomposite surface, offering a high concentra- slight redshift (≈ 6 nm) as compared to the free porphy-
tion of 4-Nitrophenol near to the Por-GO surfaces and thus rin, whereas the two Q-bands get red-shifted, shortened

13
Journal of Materials Science: Materials in Electronics

Scheme 4  Mechanism of cata-
lytic reduction of 4-Nitrophenol
to 4-Aminophenol in presence
of ­NaBH4 and Por-GO compos-
ite as catalyst

and flattened to the baseline. These results corroborate Acknowledgments  We would like to acknowledge the Council of Sci-
not only the linkage of porphyrin with GO sheets but also entific and Industrial Research (CSIR), New Delhi under Grant No.
09/100(0184)/2015-EMR-I for financial assistance. We thank SAIF,
indicate that the attachment of the porphyrin moiety has Punjab University Chandigarh for powder X-ray diffraction study,
perturbed both the ground electronic state of the GO as Indian Institute of Technology Mandi for SEM and TEM studies.
well as porphyrin. In the PXRD pattern of Por-GO nano-
composite, a comparatively broader peak is observed at
12.1°(2θ) with a d-spacing of 7.52 Ǻ indicating the inter-
layer spacing was maintained due to the intercalation of References
porphyrin and the successful formation of Por-GO nano-
composite. The PXRD profile confirmed the amorphous 1. J. Yao, Y. Sun, M. Yang, Y. Duan, J. Mater. Chem. 22, 14313
and nanoporous character of the cross-linked nanocompos- (2012)
2. Y. Zhu, S. Murali, W. Cai, X. Li, J.W. Suk, J.R. Potts, R.S. Ruoff,
ite. The morphology and topography of the nanocompos- Adv. Mater. 22, 3906 (2010)
ite appear to have stacked lamellar structure with a large 3. G. Cardenas-Jiron, M. Borges-Martinez, E. Sikorski, T. Baruah,
number of cavities and free spaces as evidenced by SEM. J. Phys. Chem. C 121, 4859 (2017)
The graphene sheet surfaces in the nanocomposite appear 4. M. Kumar, H. Khajuria, H.N. Sheikh, J. Mater. Sci. 28, 9423
(2017)
somewhat more wrinkled and rough compared to the neat 5. A. Wang et al., Inorg. Chem. Front. 3, 296 (2016)
GO sheets which can be attributed to the covalent linkage 6. S. Stankovich et al., Nature 442, 282 (2006)
of porphyrin to the GO sheet. The morphological features 7. B.Y. Xu et al., Adv. Mater. 21, 1275 (2009)
and extent of covalent linking are further supported by 8. P.V. Kamat, J. Phys. Chem. Lett. 2, 242 (2011)
9. B. Zhang et al., J. Polym. Sci. 50(2), 378 (2012)
TEM analysis. HR-TEM images of RGO reveal a wrinkled, 10. A. Kumar, P. Kumar, S. Paul, S.L. Jain, App. Surf. Sci. 386, 103–
curled and corrugated sheet typical of a few layered struc- 114 (2016)
tures. The existence of small nanoparticles (dark spot) 11. P. Kumar, A. Kumar, C. Joshi, R. Singh, S. Saran, S.L. Jain, RSC
distributed uniformly on the wrinkled graphene sheets of Adv. 5, 42414–42421 (2015)
12. S. Katoch, G.D. Bajju, G. Devi, A. Ahmed, J. Therm. Anal. Calo-
Por-GO nanocomposite further supports the functionaliza- rim. 130, 2157 (2017)
tion of GO by the covalent bonding between N­ H2SA-NiPor 13. P. Dechan, G.D. Bajju, P. Sood, U.A. Dar, J. Mol. Str. 1183, 87–99
and GO. (2019)
After the successful synthesis and characterization, the 14. R. Yamuna, S. Ramakrishnan, K. Dhara, R. Devi, N.K. Kothurkar,
E. Kirubha, P.K. Palanisamy, J. Nanopart. Res. 15, 1399–1408
Por-GO nanocomposite was further tested for the photo- (2013)
catalytic reduction of 4-Nitrophenol as the representa- 15. C.M. Drain, A. Varotto, I. Radivojevic, Chem. Rev. 109, 1630
tive substrate under the visible light irradiation using (2009)
sodium borohydride as reductant. The reduction reaction 16. J.E. Redman, N. Feeder, S.J. Teat, J.K.M. Sanders, Inorg. Chem.
40, 2486 (2001)
of 4-Nitrophenol to 4-Aminophenol catalyzed by Por-GO 17. Y. Watanabe, H. Fujii, in Metal-oxo and metal-peroxo species
nanocomposite results in the color change of the substrate in catalytic oxidations, structure and bonding, vol. 97, ed. by B.
solution from dark yellow to colourless, associated with a Meunier (Springer, Berlin, 2000), p. 61
characteristic hypsochromic shift in the absorption peak 18. Z.B. Liu, Y.F. Xu, X.Y. Zhang, X.L. Zhang, Y.S. Chen, J.G. Tian,
J. Phys. Chem. B 113, 9681–9686 (2009)
from 400  nm to 300  nm, indicating the conversion of 19. M.B.M. Krishna, N. Venkatramaiah, R. Venkatesan, D.N. Rao, J.
4-Nitrophenolate to 4-Aminophenol. Therefore, the Por- Mater. Chem. 22, 3059–3068 (2012)
GO nanocomposites serve as an efficient nanocatalyst for 20. M.B.M. Krishna, V.P. Kumar, N. Venkatramaiah, R. Venkatesan,
the catalytic reduction of organic pollutants in the treat- D.N. Rao, Appl. Phys. Lett. 98, 081103–081106 (2011)
21. W.S. Hummers Jr., R.E. Offeman, J. Am. Chem. Soc. 80(6), 1339
ment of contaminated water and help reduction in water (1958)
pollution.

13
Journal of Materials Science: Materials in Electronics

22. D.C. Marcano, D.V. Kosynkin, J.M. Berlin, A. Sinitskii, Z. Sun, 43. N.A. Kumar, H.J. Choi, Y.R. Shin, D.W. Chang, L. Dai, J.B. Baek,
A. Slesarev, L.B. Alemany, W. Lu, J.M. Tour, ACS Nano 4, 4806 ACS Nano 6, 1715–1723 (2012)
(2010) 44. S. Panigrahi et al., J. Phys. Chem. C 111, 4596 (2007)
23. A. Ahmed, Ashu, G. Devi, S. Katoch, G.D. Bajju, Ind. J. Heter. 45. N. Comisso et al., Electrochem. Commun. 25, 91 (2012)
Chem. 27, 255–264 (2017) 46. X.F. Zhang, X.Y. Zhu, J.J. Feng, A.J. Wang, Appl. Surf. Sci. 428,
24. H.M. Ahn, J.M. Bae, M.J. Kim, K.H. Bok, H.Y. Jeong, S.J. Lee, 798 (2018)
C. Kim, Chem. 23, 11969–11976 (2017) 47. K.B. Narayanan, N. Sakthivel, J. Hazard. Mater. 189, 519 (2011)
25. G.D. Bajju, Ashu, A. Ahmed, G. Devi, BMC Chem. 13, 15 (2019) 48. D. Makovec, M. Sajko, A. Selisnik, M. Drofenik, Mater. Chem.
26. G.D. Bajju, S. Katoch, G. Devi, S. Kundan, M. Bhagat, Main Phys. 129, 83 (2011)
Group Met. Chem. 39, 19–29 (2016) 49. T.L. Lai, K.F. Yong, J.W. Yu, J.H. Chen, Y.Y. Shu, C.B. Wang, J.
27. S. Sakthinathan, S. Kubendhiran, S.M. Chen, K. Manibalan, M. Hazard. Mater. 185, 366 (2011)
Govindasamy, P. Tamizhdurai, Electroanalysis 28, 1 (2016) 50. P. Babji, V.L. Rao, Int. J. Chem. Studies 4, 123 (2016)
28. Y. Xu, H. Bai, G. Lu, C. Li, G. Shi, J. Am. Chem. Soc. 130, 5856 51. Y. Liu, Y.Y. Zhang, Q.W. Kou, Y. Chen, D.L. Han, D.D. Wang,
(2008) Z.Y. Lu, L. Chen, J.H. Yang, S. Xing, RSC Adv. 8, 2209 (2018)
29. A. Bakour, M. Baitoul, O. Bajjou, F. Massuyeau, E. Faul- 52. M.T. Shah, A. Balouch, A.A. Pathan, A.M. Mahar, S. Sabir, R.
ques, Mater. Res. Express. 4, 025031 (2017). https ​ : //doi. Khattak, A.A. Umar, Microsyst. Technol. 23(12), 5745–5758
org/10.1088/2053-1591/aa5ad​4 (2017)
30. O. Bajjou, A. Bakour, M. Khenfouch, M. Baitoul, E. Faulques, M. 53. E. Marais, T. Nyokong, J. Hazard. Mater. 152, 293 (2008)
Maaza, Synth. Met. 221, 247–252 (2016) 54. X. Zhao, Y. Wang, W.H. Feng, H.T. Lei, J. Li, RSC Adv. 7, 52738
31. O. Bajjou, A. Bakour, M. Khenfouch, M. Baitoul, B.M. Mothudi, (2017)
M. Maaza, E. Faulques, J. Mater. Sci. 29, 8594–8600 (2018) 55. Y. Chen, Y. Zhang, Q. Kou et al., Nanomaterials 8, 353 (2018)
32. Z. Liu, Y. Wang, X. Zhang, Y. Xu, Y. Chen, J. Tian, Appl. Phys. 56. M. Gangarapu, S. Sarangapany, K.K. Veerabhali, S.P. Devi-
Lett. 94, 021902 (2009) priya, V.B.R. Arava, J. Clust. Sci. https​://doi.org/10.1007/s1087​
33. T. Umeyama et al., J. Phys. Chem. C 111, 11484 (2007) 6-017-1280-3
34. N. Karousis, A.S.D. Sandanayaka, T. Hasobe, S.P. Economopou- 57. R. Krishnaa et al., Int. J. Hyd. Energy 40, 4996 (2015)
los, E. Sarantopoulos, N. Tagmatarchis, J. Mater. Chem. 21, 109 58. R. Krishnaa, D.M. Fernandesb, V.F. Domingosc, E.S. Ribeirod,
(2011) J.C. Gild, E. Titus, RSC Adv. https​://doi.org/10.1039/c5ra0​5523g​
35. Y. Xu, L. Zhao, H. Bai, W. Hong, C. Li, G. Shi, J. Am. Chem. Soc. 59. T. Aditya, A. Pal, T. Pal, Chem. Commun. 51, 9410–9431 (2015)
131, 13490 (2009) 60. Z.D. Pozun, S.E. Rodenbush, E. Keller, K. Tran, W. Tang, K.J.
36. M. Jahan, Q. Bao, K.P. Loh, J. Am. Chem. Soc. 134, 6707 (2012) Stevenson et al., J. Phys. Chem. C 117, 7598–7604 (2013)
37. D. Geng, S. Yang, Y. Zhang, J. Yang, J. Liu, R. Li, T.K. Sham, X. 61. P. Zhao, X. Feng, D. Huang, D. Astruc, Coord. Chem. Rev. 287,
Sun, S. Ye, S. Knights, Appl. Surf. Sci. 257, 9193–9198 (2011) 114 (2015)
38. N.A. Kumar et al., J. Mater. Chem. A 5, 13204 (2017) 62. T. Aditya, A. Pal, T. Pal, Chem. Commun. 00, 1 (2012)
39. R. Krishna, E. Titus, L.C. Costa, J.C. Menezes, M.R. Correia, S. 63. R. Krishna, E. Titus, O. Okhay, J.C. Gil, J. Ventura, R.E. Venkata,
Pinto, J. Mater. Chem. 22, 10457–10459 (2012) Int. J. Electrochem. Sci. 9, 4054 (2014)
40. B. Yao, C. Li, J. Ma, G. Shi, Phys. Chem. Chem. Phys. 17, 19538 64. S. Bai, X. Shen, G. Zhu, M. Li, K. Chen, A.C.S. Appl, Mater.
(2015) Interfaces 4, 2378 (2012)
41. S. Kulbendhiran, S. Saktinathan, S.-M. Chen, P. Tamizhdurai, J.
Colloid Interface Sci. 497, 207 (2017) Publisher’s Note Springer Nature remains neutral with regard to
42. R. Rahimi, R. Zare-Dorabei, A. Koohi, S. Zargari, Proceeding of jurisdictional claims in published maps and institutional affiliations.
the 18th International Electronic Conference on Synthetic Organic
Chemistry, Lugo (2014)

13

You might also like