You are on page 1of 20

Res Chem Intermed

DOI 10.1007/s11164-017-3161-7

A facile green synthesis of silver nanoparticle-decorated


hydroxyapatite for efficient catalytic activity towards
4-nitrophenol reduction

Tushar Kanti Das1 • Sayan Ganguly1 • Poushali Bhawal1 •

Subhadip Mondal1 • Narayan Ch Das1

Received: 15 July 2017 / Accepted: 25 September 2017


 Springer Science+Business Media B.V. 2017

Abstract Silver nanoparticles (AgNPs) were gradually grown on a hydroxyapatite


surface via a green self-reducing approach. Hydroxyapatite (HAP) was functionalized by
dopamine and subsequently self-polymerized onto the HAP surface. Surface-adhered
polydopamine acted as a self-reducing agent for the Ag? ion; thus the process is con-
sidered to be green. The nanoparticles were evaluated by UV–visible spectroscopy,
X-ray diffraction, scanning electron microscopy, transmission electron microscopy, and
energy-dispersive X-ray spectroscopy. The characterization data revealed that the
AgNPs were spherical and crystalline, with an average size of 14.79 nm. The AgNP-
decorated HAP was applied as a heterogeneous catalyst for the reduction of 4-nitro-
phenol (4-NP), which obeyed first-order reduction kinetics. Our reported catalyst can
thus be used as an eco-friendly, efficient, cost-effective, and easy-to-fabricate hetero-
geneous catalyst system with potential use in industrial applications.

& Narayan Ch Das


ncdas@rtc.iitkgp.ernet.in
1
Rubber Technology Centre, Indian Institute of Technology, Kharagpur 721302, India

123
T. K. Das et al.

Graphical Abstract

Keywords Silver nanoparticles  Hydroxyapatite  Polydopamine  Heterogeneous


catalyst  4-Nitrophenol  First-order reduction

Introduction

Over the last few decades, accelerated industrial growth has led to the release of an
enormous amount of wastewater containing dyes, phenol and its derivatives, and
other organic pollutants, resulting in increased environmental hazards. This water
must be rehabilitated before it is discharged into the main water systems [1–3]. The
complete eradication of organic pollutants from wastewater is essential, and has
become a subject of extensive research. Several methods have been explored for the
partial or complete eradication of these pollutants, including coagulation, adsorp-
tion, reverse osmosis, biodegradation, and chemical or photochemical degradation
[4, 5]. With the exception of chemical or biodegradation, these methods are
somewhat ineffective, and further treatment is needed, which increases the cost. In
the case of biodegradation, problems arise not only in the antagonistic effect of
highly concentrated effluents towards microorganisms, but also in the resistence to
both chemical and physical destruction. Chemically stabilized nanoparticles have
recently been employed for protracted wastewater treatment [6–9].
The introduction of the principles of green chemistry into nano-science marked
the beginning of a universal trend in the world of science and technology. Tuning
the size of nanoparticles to control their efficiency is important for applications in
such fields as biomedicine, cosmetics, electronics, pharmaceuticals, and energy and
the environment [10–12]. Therefore, maintaining the size of nanoparticles during
their synthesis via green chemistry has become a challenge for nano-science

123
A facile green synthesis of silver nanoparticle-decorated…

researchers. On the other hand, the ability to stabilize while simultaneously gaining
activity from nanoparticles depends upon the choice of supporting materials. These
materials, through their specific interaction with the nanoparticles, strongly
influence the above-mentioned properties [13, 14].
Hydroxyapatite (HAP), having the formula Ca10(PO4)6(OH)2, is an inorganic
compound that is an integral component of bones and teeth [15, 16]. From an
environmental point of view, it is a benign material for living organisms. Therefore,
it has garnered interest among researchers for potential use in biomedical
applications such as bone fillers [17], scaffolds for bone tissue [18], a promoter
of soft tissue repair [19], drug loading and delivery systems [20], and a solid phase
for the separation of biomolecules via column chromatography [21]. High surface
energy and nano-range HAP particles are prone to aggregation and can be
incompatible with both inorganic and organic matrices [22, 23]. To overcome this
problem, numerous surface modification strategies have been proposed to enhance
adhesion with other materials [24, 25].
Researchers have recently found that 3,4-dihydroxy-L-phenylalanine (dopamine)
plays a significant role in the adhesive properties of mussels, enabling the formation
of a potent coating for adherence to rocks or boats [26–28]. Because of its efficacy
as an adhesive and its viability in the synthesis of PDA by simple oxidative
polymerization, dopamine (DOPA) is used extensively for the coating or
modification of a variety of substrates, including metals and metal-oxides [29],
carbonaceous materials [30], and polymers [31]. In the past, numerous phenol
derivatives have been employed in the synthesis of AgNPs or gold nanoparticles
(AuNPs), serving a dual role as a reductant and stabilizer [32, 33]. DOPA has
garnered significant attention because it bears amine and catechol functional groups.
The catechol functional groups, presumably present after PDA coating, are readily
oxidized and can effectively reduce metal salts from their higher oxidation state to a
metallic state [27, 34]. Thus, DOPA can undergo spontaneous oxidative polymer-
ization (PDA) under mild conditions to coat various substrates, and it also has the
ability to reduce and stabilize metallic nanoparticles while increasing the
compatibility between organic and inorganic materials [27, 35].
With this in mind, the aim of the present work was to prepare efficient catalytic
AgNP-decorated PDA-coated HAP nanocomposites, where the PDA macro-
moleculer chains serve the purpose reducing agent as well as stabilizer for the
fabrication of in situ AgNPs over HAP surfaces. The catalytic nanocomposites
developed herein showed high efficiency towards the catalytic reduction of phenolic
derivatives in the presence of sodium borohydride (NaBH4). To analyze the
catalytic efficiency of the nanocomposites, we examined the catalytic reaction on
the basis of pseudo-first-order kinetics. The high stability of the nanocomposites
was also assessed by performing several cycles of the same catalytic reduction.
These highly effective and reusable catalytic nanocomposites can thus be used as
heterogeneous catalysts for the purification of water, as well as other purification
and separation, within the framework of green chemistry.

123
T. K. Das et al.

Experimental

Materials

Analytical-grade calcium nitrate tetra-hydrate [Ca(NO3)24H2O], ammonium


hydrogen phosphate [(NH4)2HPO4], dopamine hydrochloride [3,4-dihydrox-
yphenethylamine hydrochloride], Tris base [2-amino-2-(hydroxymethyl)-1,3-
propanediol], p-nitro phenol(4-NP), o-nitro phenol (2-NP), m-nitro phenol (3-NP)
and sodium borohydride (NaBH4) were obtained from Sigma-Aldrich (St. Louis,
MO, USA) and were used as received without further purification. Silver nitrate
[AgNO3], hydrogen chloride (HCl), Congo red (CR) and ammonium (NH3)
(aqueous solution) were purchased from Merck (Darmstadt, Germany). Millipore
water was used in all of the experiments.

Synthesis of hydroxyapatite (HAP)

HAP nanorods were synthesized via a simple wet chemical process that relies on the
precipitation of forerunner from aqueous solution [36, 37]. A brief description of
HAP nanorod synthesis is as follows: First, Ca(NO3)24H2O and (NH4)2HPO4 were
separately dissolved in water while maintaining the atomic ratio of calcium and
phosphate at 1.67. The pH of the solution was adjusted to 10 using aqueous NH3
solution. Next, (NH4)2HPO4 solution was slowly added to Ca(NO3)24H2O solution,
and the temperature was maintained at 90 C using an oil bath. The mixture was
stirred for 5 h at a constant temperature of 90 C. Finally, the suspension was
thoroughly centrifuged with water several times to remove unreacted Ca(NO3)2-
4H2O and (NH4)2HPO4, and some of the precipitate was left in the aqueous phase
for further modification. For the characterization of HAP, the remainder of the
solution was dried at 60 C in a vacuum oven overnight.

Synthesis of PDA-coated HAP (HAPM)

For this step, 0.2 g of nano-HAP was dispersed in 100 ml Tris base [10 mM,
maintaining pH = 8.5 by adding HCl (aq)] solution via ultrasonication for 30 min.
DOPA hydrochloride was subsequently introduced to this mixture until a DOPA
concentration of 2 mg/ml was achieved. Following the addition of DOPA, the
solution gradually changed from white to black, as shown in Fig. 1, indicating the
in situ polymerization of DOPA to PDA. The modification reaction was performed
for 24 h at room temperature with continuous stirring. PDA-coated HAP (HAPM)
was then centrifuged with water three times to remove unreacted DOPA, and the
black sample was dried at 60 C in a vacuum oven overnight prior to its subsequent
reaction.

123
A facile green synthesis of silver nanoparticle-decorated…

Fig. 1 Digital image of the conversion from synthesized nano-hydroxyapatite to AgNP-adsorbed


polyDOPA-coated hydroxyapatite in aqueous solution (RT room temperature)

Synthesis of AgNP-decorated, PDA-coated HAP nanocomposite (HAPX)

In a typical preparation of AgNP-immobilized HAPM, 100 mg HAPM was


suspended in an aqueous solution (100 ml) of silver nitrate (AgNO3), and the
mixture was stirred for 24 h. We prepared solutions with 1, 2, 3, 4 and 5 mM of Ag
(NO3) before the addition of 100 mg HAPM. The prepared AgNP-immobilized
HAPM nanocomposites were later denoted as HAPX, where X denotes AgNO3
concentration. AgNP-modified samples were subsequently separated by centrifu-
gation and washed with water to remove unreacted AgNO3. Finally, the samples
were dried at 60 C in a vacuum oven overnight prior to further use. Figure 1 shows
a digital image of three stages during the conversion of product, starting from the
aqueous dispersion of HAP to HAPX.

Catalytic reduction of nitrophenols in the presence of NaBH4


by nanocomposites

The reduction of 4-NP in the presence of excess NaBH4 under the action of the
prepared nano-catalysts was carried out as a model reaction to examine their
catalytic activity. The catalytic reduction of 4-NP to p-aminophenol (4-AP) was
conducted in a 100-ml breaker containing an aqueous solution of 4-NP (30 ml,
0.1 mM). A freshly prepared aqueous solution of NaBH4 (0.2 M, 3 mL) was added
to the solution, and the light yellow color of the solution instantly gave way to a
bright yellow color. Next, 2 mg of the prepared catalyst was immediately
introduced into the stirred solution at room temperature, and time-dependent
absorption spectra of aliquots withdrawn at 1-min intervals were recorded using an
ultraviolet–visible (UV–Vis) spectrometer over a scanning range of 200–800 nm.
The kinetic study of the reaction was based on the characteristic absorption peak of
4-NP at 400 nm (characteristic peak of p-nitro-phenolate ion) as a function of time.
To examine the reusability of the prepared catalyst, it was removed by
centrifugation upon completion of the reduction process. The sample was then
washed with water and ethanol several times, and was subsequently dried at 60 C
in a vacuum oven prior to the next cycle. The catalytic reduction of 2-NP (0.1 mM)
and 3-NP (0.1 mM) in the presence of the HAP5 nanocomposite was also performed
under conditions similar to those for 4-NP.

123
T. K. Das et al.

Catalytic degradation of Congo red (CR) in the presence of NaBH4


by nanocomposites

The catalytic activity of the HAP5 nanocomposites was also evaluated by


monitoring the degradation of an aqueous solution of Congo red (CR) in the
presence of NaBH4. For the degradation studies, 10 mg of HAP5 nanocomposites
was added to a 30-ml CR solution (20 ppm) in a 100-ml beaker. Next, 5 ml of a
freshly prepared aqueous solution of NaBH4 (0.2 M) was introduced to the mixture,
which was continuously stirred with a magnetic stirrer at room temperature. Then,
aliquots were withdrawn from this mixture in 1-min intervals, and the concentration
of the CR solution was measured by UV–Vis spectroscopy to construct time-
dependent UV–Vis absorption spectra.

Characterization techniques

The X-ray diffraction (XRD) data derived from the prepared HAP and its modified
samples were recorded on an X’Pert PRO system (PANalytical B.V., Almelo, the
Netherlands) using Ni-filtered Cu Ka radiation (k = 1.5418 Å) and a scanning rate
of 0.005 (2h/s). Fourier transform infrared (FTIR) spectra derived from these
samples were recorded on an FTIR spectrometer (Spectrum Two; PerkinElmer,
Waltham, MA, USA) using KBr pellets in a frequency range of 600–4000 cm-1 and
a resolution of 4 cm-1. For high-resolution transmission electron microscopy
(HRTEM) imaging, samples were dispersed in water at the desired concentration by
ultrasonication for 30 min, followed by drop-casting onto a carbon-coated Cu grid
using a graduated micropipette. The samples were then dried under a vacuum at
50 C for 1 day. Finally, the samples were analyzed using HRTEM (JEM-2100;
JEOL, Tokyo, Japan), with a tilt angle of 24, operated at 200 kV, filament LaB6).
For the field emission scanning electron microscopy (FESEM) studies, 1 mg of each
sample was sonicated in 20 ml of water for 30 min, drop-cast on a silicon wafer,
and dried under a vacuum at 50 C for 1 day. Before FESEM imaging, the drop-cast
samples were gold-coated in a sputter coater. These samples were examined at
different levels of magnification by FESEM (Merlin with tungsten; Carl Zeiss,
Germany), with stimulation voltage set to 15 kV, to gain information about the
morphology of the samples. Energy-dispersive X-ray (EDX) and mapping of the
silver-modified samples was also accomplished by FESEM. The reaction kinetics of
the nitrophenol reduction and the degradation of CR in the presence of different-
concentration AgNP-immobilized hydroxyapatite samples was monitored by UV–
Vis spectroscopy (UV-1601, Shimadzu, Japan).

Results and discussion

Correlation of the reaction rates of different catalysts requires an orderly and


consistent strategy for synthesizing the catalyst, along with a constant environment
during the catalytic reaction. To determine the catalytic efficiency of the newly

123
A facile green synthesis of silver nanoparticle-decorated…

synthesized catalysts, reduction of 4-NP in the presence of excess NaBH4 was


studied as a model pseudo-first-order reaction. The mechanism for the reduction of
4-NP to 4-AP is complicated by factors including the reliability of electron transfer
to nitro compounds and the accessibility of hydrogen atoms [38]. The catalytic
efficiency of the reduction depends mainly on the electron transmission power of the
catalysts, as shown in Fig. 2. The reduction of 4-NP can be physically monitored by
the disappearance of the bright yellow color and the gradual decrease in the
absorbance peak at 400 nm (characteristic peak of p-nitrophenolate ion) in UV–Vis
spectrophotometry.

Morphological analysis of prepared samples

Morphological analysis is important for the study of the organizational and


physiological state of materials. Thus, it can generally be applied for the qualitative
and, in some cases, quantitative measurement of materials. HRTEM and FESEM
along with EDX were performed to explore the organizational and physiological
state of the various prepared samples. Figure 3 shows typical HRTEM images of
HAP, HAPM and HAPX. Rod-like HAP crystals 100–200 nm in length are shown
in Fig. 3d, which is consistent with the results of FESEM. The diameter of these
rod-like crystals is very small relative to their length. These HRTEM images
confirm the rod-like shape of the synthesized HAP crystals at the nanoscale, which
is also supported by XRD studies (hexagonal crystal). As the crystals exist in the
nano-range without the addition of stabilizer, they tend to produce clusters [39].

Fig. 2 Proposed mechanism of reduction of 4-NP to 4-AP by catalytic sample

123
T. K. Das et al.

Fig. 3 a HRTEM image of synthesized rod-like nano-HAP crystal, b PDA-coated HAP (inset shows thin
coating of PDA over HAP), c AgNPs decorated on coated HAP (HAP3; inset: selected area [electron]
diffraction [SEAD] pattern), d magnified view of (a) to determine the length of rods

Following the in situ polymerization of DOPA on the surface of the HAP crystal, the
propensity towards aggregation is reduced, and this is clearly demonstrated in
Fig. 3b. The inset of these figures displays the distinct coating of PDA on the
surface of the HAP crystal. Figure 3c shows an HRTEM image of the absorbed
AgNPs, which are randomly distributed on the surface of HAPM and have an
estimated particle size of roughly 20 nm, a value which is more or less consistent
with the average size obtained from FESEM images using IMAGEJ software
(discussed below). The inset of Fig. 3c shows a selected area of the diffraction
pattern of AgNP-decorated HAPM. The bright concentric ring implies that the
nanoparticles are polycrystalline.
The morphology of the samples was also analyzed by FESEM and is depicted in
Fig. 4. The rod-like but aggregated structure of HAP is clearly shown in Fig. 4a and

Fig. 4 a FESEM image of synthesized rod-like nano-HAP crystal, b PDA-coated HAP, c AgNPs
decorated on coated HAP (HAP3)

123
A facile green synthesis of silver nanoparticle-decorated…

is in accord with previous results [40]. After modification of the HAP surface with
PDA, it became smooth, indicating the formation of a thin coating of PDA, as
depicted in Fig. 4b. Figure 4c reveals that nanoparticles reduced by PDA are
randomly distributed over HAPM. A few AgNPs decorated on the surface of HAP
(shown in Fig. 4c) are marked by circles, and their corresponding diameters are
given below the circles. The size of these AgNPs was measured from the FESEM
image, and the corresponding histogram is displayed in Fig. 5. The average
diameter of the AgNPs obtained from the histogram, after fitting with a Gaussian
distribution curve, was 15 nm. The chemical composition was evaluated by the
EDX spectrum obtained from the FESEM analysis, as shown in Fig. 6, and the
corresponding mass percentage and atomic percentage are given in Table 1. The
spectrum revealed the mass ratio as well as the atomic ratio of the catalyst surface.
The C:O:Ag ratio was calculated to be 4.00:18.38:47.86 (wt%) and
12.11:68.83:16.15 (at%) on the exposed surface. The C and O originate from the
PDA adhered to the HAP surface. Hence, the presence of C, O and Ag obtained
from the EDX surface scan supports the successful synthesis of AgNP-decorated
PDA-coated hydroxyapatite via a green method.

Spectroscopic analysis of prepared samples

Spectroscopic analysis provides information about the chemical structure of


materials as well as changes to their chemical structure following modification via
chemical reaction. We carried out an FTIR and XRD study to detect changes in the
structure of the modified HAP. FTIR spectra of HAP, HAPM and various HAPM
with absorbed AgNPs are compared in Fig. 7. To obtain a good comparison between
them, we highlighted a small region of the FTIR spectra. This analysis was
performed to ascertain the occurrence of any chemical reaction between the
inorganic and organic (polymer) phases, and it was used to determine structural
changes during modification. The synthesized HAP shows characteristic peaks at
1040 and 1090 cm-1 (stretching of P–O bond), indicating the presence of PO4
groups. This is very similar to those observed for naturally occurring hydroxyapatite
[41]. The small weak peak at 962 cm-1 is attributed to the presence of PO4 groups.
The peak at 627 cm-1 arises from the out-of-plane bending mode vibration of –OH

Fig. 5 AgNP size distribution


histogram obtained from Fig. 4c

123
T. K. Das et al.

Fig. 6 EDX spectrum of AgNPs decorated on coated HAP (HAP5)

Table 1 Mass percentage and


Elements Wt% At%
atomic percentage of HAPX for
each element obtained from
CK 4.00 12.11
EDX spectrum
NK 1.71 4.46
OK 18.38 41.81
AgL 47.86 16.15
CaK 28.05 25.48

Fig. 7 FTIR spectra of HAP, PDA-coated HAP, and various AgNPs and PDA-coated HAP
nanocomposite

123
A facile green synthesis of silver nanoparticle-decorated…

groups [42]. The peaks at 627, 567 and 481 cm-1 are ascribed to the bending mode
of the vibration of P–O bonds. A broad but small peak around 830 cm-1 is
attributed to stretching of the P–OH groups [43]. This FTIR analysis implies that
HAP was successfully synthesized. Because a thin coating of PDA was formed on
the surface of nano-HAP, the FTIR spectra of HAPM do not show a noticeable
change. Nevertheless, the stretching band of HAPM was shifted around 7 cm-1
compared to that of the synthesized HAP, which clearly indicates a strong
interaction between HAP and PDA [44]. The spectra of each HAPM after
immobilization with AgNPs are almost indistinguishable from those of HAPM,
indicating no significant change to the internal structure of HAPM.
The XRD patterns derived from HAP, HAPM and various AgNPs adsorbed on
HAPM are shown in Fig. 8. The XRD pattern of synthesized powder nano-HAP
reveals the existence of a crystalline phase, and from the listed database it was found
to be consistent with JCPDS (No. 9-04323). The main characteristic planes for
nanometer-size HAP are (002), (210), (211), (112), (300), (310), (222), (213) and
(004) [37, 45], and the corresponding angles (2h) are 26, 29, 32.15, 33.1, 34.65,
40.2, 46.31, 49.68 and 53, respectively, as shown in Fig. 8. The characteristic
peaks are analogous to hexagonal crystalline HAP [46, 47]. The mean size of the
synthesized nano-HAP crystallite was determined by the Debye–Scherrer equation
[48]:
thkl ¼ 0:9k=B cosðhhkl Þ ð1Þ
where t(hkl) is the crystallite size, k is the wavelength of the X-ray beam, B is the full
width at half maximum (FWHM) of the characteristic peak at maximum intensity,
and h(hkl) is the diffraction angle for the (hkl) plane. The characteristic (002)
reflection peak was taken into account to measure the size of the crystallite, and it
was found to have an average value of approximately 36 nm. The XRD pattern of
HAPM is similar to that of nano-HAP, which signifies that no alternation of the
crystalline phase of HAP has taken place following modification via the in situ

Fig. 8 XRD of HAP, PDA-coated HAP, and various AgNPs and PDA-coated HAP

123
T. K. Das et al.

polymerization of DOPA. The XRD diffraction patterns of each of the AgNP-


adsorbed HAPM nanocomposites showed characteristic peaks attributed to AgNPs,
along with a characteristic peak of HAP. Figure 8 also reveals that the characteristic
2h peak at 36.67, 48.19, 64.1 and 76.88 and the corresponding reflection plane
(111), (200), (220) and (311) are consistent with the face-centered cubic (FCC)
structure of metallic AgNPs adsorbed on the surface of HAPM [49, 50]. Hence, the
crystallinity of HAP and the AgNPs following modification are distinct.

Mechanism for stabilization of AgNPs decorated on in situ polymerized


DOPA-modified nano-HAP

Figure 9 depicts the reaction mechanism for the coating of synthesized rod-like
nano-HAP via the in situ polymerization of DOPA, as well as the stabilization of
AgNPs. Nano-HAP bearing surface –OH groups can undergo reaction with either
the –NH2 or –OH groups of DOPA to covalently attach to the nanorod surface.
Next, DOPA undergoes oxidative polymerization in the presence of Tris base (at
pH = 8.5, and room temperature) to yield PDA on the surface of HAP, as
previously described in the literature [27]. The catechol groups of PDA at the
terminal end of the polymer chains subsequently mediate the reduction of Ag? to
Ag state, while undergoing oxidation to form the corresponding quinone [51, 52].
Finally, Ag atoms assemble into AgNPs clusters, which gain stability from the
remaining catechol groups of PDA [53, 54]. In summary, PDA essentially acts as a
reductant, while stabilizing AgNPs by prohibiting their agglomeration.

Mechanism for reduction of 4-NP by the synthesized catalyst

The reduction of 4-NP in the presence of NaBH4 alone is ineffective, and if the
reaction does take place, it is slow. Thus, small particles with high surface area are
necessary to promote the reduction [55]. Based on this observation, the mechanism
for the reduction of 4-NP by the synthesized catalysts is depicted in Fig. 2. The

Fig. 9 Mechanism for coating of PDA on nano-HAP and stabilization of AgNPs by PDA

123
A facile green synthesis of silver nanoparticle-decorated…

reduction is basically divided into four distinct stages. First, adsorption of active
hydrogen atoms on the catalytically active NPs is induced by the reaction of NaBH4
and H2O. Second, adsorption of 4-NP by its -NO2 groups occurs on the remainder of
the catalytically active sites of the NPs. Next, reduction of the -NO2 groups to –NH
2 groups by the neighboring active hydrogen atoms occurs, and finally, desorption of
the produced 4-AP from the active surface of the NPs delivers the product and
liberates the catalyst for the next cycle. Among these four steps, the adsorption of
hydrogen atoms on the surface of NPs is crucial, and it is the rate-determining step
of the catalytic reduction. Hence, the overall rate of reaction depends upon the
activity of the catalyst produced.

Monitoring catalytic reductions of nitrophenols by UV–Vis spectra

After complete characterization of the AgNPs immobilized on HAPM, in view of


the notable catalytic activity of stabilized AgNPs, each AgNP-modified sample was
employed for the catalytic reduction of 4-NP in the presence of NaBH4. The
successful conversion of 4-NP to 4-AP was monitored by UV–Vis spectroscopy.
Specifically, 4-NP shows an absorption peak at 317 nm (not shown here), but in the
presence of NaBH4, the peak is shifted to 400 nm due to the formation of the p-
nitro-phenolate ion. In the presence of a very small amount of the catalytic sample,
the characteristic absorption peak height at 400 nm decreases, and a new peak
simultaneously appears at 300 nm, implying the formation of 4-AP, as shown in
Fig. 10. For each prepared nano-catalyst, the catalytic reduction is completed within
10 min, and as we go from HAP1 towards HAP5, the time required for completion

Fig. 10 Time-dependent UV–Vis spectra of HAP 1, HAP2, HAP3, HAP4 and HAP5 for the catalytic
reduction of 4-NP

123
T. K. Das et al.

of the reaction is reduced to 5 min. However, in each case, an aliquot is withdrawn


from the reaction mixture after 10 min for UV–Vis measurement. Based on the
absorption observed at 400 nm by UV–Vis spectroscopy, we applied a pseudo-first-
order reaction kinetic (as shown below) analysis to determine the rate constant of
the reaction [56]:
ln½ðAt =A0 Þ ¼ kt ð2Þ
where At and A0 represent the absorbance (concentration) at time t and t = 0,
respectively; k = rate constant (s-1) and t = time.
The plot of (At/A0) versus time to obtain information about the activity of the
catalyst in the reduction of 4-NP is shown in Fig. 11a. After plotting of ln (At/A0)
versus time to determine the rate constant for each catalytic reaction, as shown in
Fig. 11b, the rate constant and the regression coefficient for each reaction were
obtained, and these values are summarized in Table 2. These data show that
increasing the concentration of the AgNO3 precursor gives rise to a rate constant
increase, which indicates higher efficacy in the reduction. This may be due to the
generation of more nanoparticles on the surface of HAPM by PDA, leading to
higher catalytic activity. The generation of more AgNPs means that more
nanoparticles are available to absorb active hydrogen and 4-NP, resulting in
increased nano-catalyst efficiency [57]. The rate constant obtained is comparable to
or significantly higher than results previously reported under the same environment
[6, 7, 54, 58, 59]. A comparison of the rate constant with previously reported results
is shown in Table 3.
For comparison of its catalytic activity, we also studied the catalytic efficacy of
highly efficient HAP5 nanocomposites towards reduction of 2-nitrophenol (2-NP)
and 3-nitrophenol (3-NP). Figure 12a, b shows time-dependent UV–Vis spectra of
the catalytic reduction of 2-NP and 3-NP by NaBH4 in the presence of the HAP5
nanocomposite. In the case of 2-NP, the characteristic absorption peak at 416 nm
(resulting from the formation of the 2-nitrophenolate ion in the presence of NaBH4)
exhibits a gradual decrease in absorption, and a new peak simultaneously appears at

Fig. 11 a Plot of (At/A0) versus reaction time (t) and b plot of ln (At/A0) versus reaction time (t) for each
catalytic sample (where At = absorbance at time t and A0 = initial absorbance)

123
A facile green synthesis of silver nanoparticle-decorated…

Table 2 Obtained rate constant


Sample Rate constant (s-1) R2 value for linear fitting
along with the regression
coefficient (R2) for each
HAP1 0.00201 0.98157
synthesized catalytic sample for
the reduction of 4-NP HAP2 0.00466 0.97659
HAP3 0.00503 0.96777
HAP4 0.00642 0.97434
HAP5 0.0073 0.98162

Table 3 Comparison of rate constant with previously reported results for the reduction of 4-NP
Sl Sample name Rate constant (s-1 9 10-3) References
no. (room temperature)

1 AgNPs/alginate hydrogel 4.06 [6]


2 Silver nano-dendrite 5.63 [7]
3 Polycaprolactone and 2-(dimethyl amino) 0.473 [54]
methacrylate copolymer/AgNPs
4 Silver nano-shell coated with cationic polystyrene 5.07 [58]
beads
5 Fe3O4 @ SiO2-Ag nanocomposite 7.67 [59]
6 In this work 7.30 –

Fig. 12 Time-dependent UV–Vis spectra of the reduction of a 2-NP and b 3-NP, and c degradation of
CR in the presence of NaBH4 by HAP5 nanocomposites

123
T. K. Das et al.

283 nm, signifying the formation of 2-aminophenol (2-AP) [60]. In the case of
3-NP, an absorption peak at 390 nm is observed (due to formation of 3-nitrophe-
nolate ion in the presence of NaBH4), and with time, the peak intensity gradually
decreases, indicating the reduction of 3-AP [61]. Because an excess of NaBH4 was
used in both catalytic reductions of 4-NP, a pseudo-first-order kinetic analysis can
be applied to resolve the rate constant based on the absorption of 2-NP and 3-NP at
416 and 390 nm, respectively. The kinetic analyses of these catalytic reactions were
studied by plotting (At/A0) versus time, as shown in Fig. 13a. The rate constants
(k) were subsequently obtained from the linear plot of ln (At/A0) versus time
(Fig. 13b), and the rate constants with their corresponding regression coefficients
are summarized in Table 4. The value of k for all catalytic reductions clearly
illustrates that the rate of the catalytic reduction of 4-NP is somewhat higher than
that of 2-NP, while the rate of reduction of 3-NP is much lower than that of both
4-NP and 2-NP. Therefore, the rate of catalytic reduction of nitrophenol pollutants
in the presence of NaBH4 by HAP5 follows the trend 4-NP [ 2-NP [ 3-NP.
Figure 10 shows that the UV–Vis absorbance spectra drift from baseline is higher
for HAP1 than for the others. Drifts in spectra have no significant effect regarding to
the degree of gradual alternation of the measurement system and feed concentration
in the whole time scale of the experimentation. In UV–Vis spectra, absorbance drift
is more significant than wavelength drift, as the wavelength is based on the
maximum of peaks which are broad in most cases. The absorbance drift can be
classified as either physical (intrinsic) drift or chemical drift. Intrinsic drift can
fluctuate on different days and depends upon the variation in parameters such as
ambient temperature and voltage. Though physical drift contributes to absorbance
drift, it is small in comparison to that of chemical drift. Chemical drift originates
mainly from the system under study rather than from equipment [62]. Chemical
absorbance drift measurement error will occur for the following reasons: (a) the
reaction in the solution under study has not reached completion at the time of
measurement; (b) the materials formed may not be stable; (c) it sometimes forms
complexes with other species present in solution. In addition, error can result from
factors associated with the instrument, such as contaminated cuvettes, variations in

Fig. 13 a Plot of (At/A0) versus reaction time (t) and b plot of ln (At/A0) versus reaction time (t) for 2-NP,
3-NP and CR by HAP5 nanocomposites

123
A facile green synthesis of silver nanoparticle-decorated…

Table 4 Comparison of rate constants for the catalytic reduction of 2-NP, 3-NP, 4-NP and CR by the
HAP5 nanocomposite
Sample Rate constant (s-1) R2 value for linear fitting

2-NP 6.05 9 10-3 0.97563


3-NP 2.56 9 10-3 0.95542
4-NP 7.30 9 10-3 0.9816
CR 5.75 9 10-3 0.9802

cell length, or improper calibration of the absorbance [63]. In the case of the UV–
Vis spectra of HAP1, the contribution of the factors leading to absorbance drift
mentioned above may be higher than for other samples (Fig. 10).

Monitoring the catalytic reduction of Congo red (CR) by UV–Vis spectra

Congo red (CR), a water-soluble azo dye with widespread use as a colorant in the
textile industry, was adopted as a model substrate for the catalytic reduction of dyes.
It displays characteristic UV–Vis absorption peaks around 498 and 340 nm due to
the presence of an azo bond (–N=N–) and a naphthalene ring, respectively [64].
Time-dependent UV–Vis absorption spectra of the catalytic degradation of CR in
the presence of NaBH4 by the HAP5 nanocomposite are shown Fig. 12c. The
characteristic absorption peaks at 498 and 340 nm decayed with time in the
presence of the HAP5 nanocomposites, indicating the degradation of CR, with
complete degradation of CR within 10 min. Based on the absorption at 498 nm, we
applied pseudo-first-order kinetic analysis to determine the rate constant. The
kinetic analysis of the catalytic degradation of CR was examined by plotting (At/A0)
versus time (Fig. 13a), and the linear correlation between ln (At/A0) and time
(Fig. 13c) implies that the catalytic degradation follows pseudo-first-order kinetics.
The rate constant derived from the slope was 5.75 9 10-3 s-1, as shown in Table 4,
which is higher than that in previously published results [65–68].

Recycling of catalysts

For practical use of heterogeneous systems, a balance must be maintained between


the activity and the reusability of the catalyst. Thus, the reusability of the nano-
catalysts was also examined in this study. Figure 14 illustrates the reusability of the
catalysts for the reduction of 4-NP with NaBH4. Catalyst activity was determined
based on the rate of reduction in comparison to the initial rate for each use. After
each catalytic reduction reaction, the nano-catalysts were recovered by centrifuga-
tion and washed repeatedly prior to their use in the next cycle, as described above.
Analysis of the results showed that the efficiency of the synthesized nano-catalyst
was almost unchanged after the fifth consecutive cycle, which suggests that there
was no loss of catalytic activity during the reaction or purification process.
Furthermore, nearly quantitative recovery of the catalyst was achieved after each

123
T. K. Das et al.

Fig. 14 Catalytic activity of the nano-catalyst (HAP5) in the reduction of 4-NP after five cycles of use

reduction. Metal nanoparticles may undergo oxidation under alkaline conditions,


leading to reduced activity [69], but the nano-catalyst produced by AgNPs in this
work are very stable and display superior catalytic activity.

Conclusion

We have described an environmentally friendly aqueous synthesis of essentially


uniform spherical AgNPs, with an average diameter of approximately 15 nm, on the
polyDOPA-coated surface of rod-like HAP, using PDA as both reductant and
stabilizer for these nanoparticles at room temperature. The stabilized AgNPs were
applied in the catalytic reduction of 4-NP in the presence of NaBH4, and results
analogous to those obtained with previously reported nano-catalysts were achieved
on the basis of the measured rate constant (k), which illustrates their catalytic
activity towards reduction. The rate constant was determined using a pseudo-first-
order kinetics model, and increased with the incremental addition of the precursor
[AgNO3 (aq.)], due to the generation of more NPs. This experimental study suggests
that the reduction of 4-NP in the presence of NaBH4 by the synthesized catalytic
nanocomposites is diffusion-controlled. On the other hand, the nano-catalyst
activity was not significantly diminished after five consecutive cycles. Hence, this
product holds potential for the rehabilitation of wastewater and the environmentally
friendly conversion of hazardous 4-NP into non-toxic 4-AP in aqueous solution
under mild conditions.

Acknowledgements This work has been financially supported by the Science and Engineering Research
Board (SERB), Department of Science and Technology (DST), Ministry of Science and Technology,
Govt. of India (ECR/2016/000048).

123
A facile green synthesis of silver nanoparticle-decorated…

References

1. N. Roostaei, F.H. Tezel, J. Environ. Manag. 70, 164 (2004)


2. M. Akçay, J. Colloid Interface Sci. 280, 304 (2004)
3. H. Wang, J. Niu, X. Long, Y. He, Ultrason. Sonochem. 15, 392 (2008)
4. A. Safavi, S. Momeni, J. Hazard. Mater. 201, 131 (2012)
5. L. Tinghui, T. Matsuura, S. Sourirajan, Ind. Eng. Chem. Prod. Res. Dev. 22, 85 (1983)
6. L. Ai, J. Jiang, Bioresour. Technol. 132, 377 (2013)
7. W. Zhang, F. Tan, W. Wang, X. Qiu, X. Qiao, J. Chen, J. Hazard. Mater. 217, 42 (2012)
8. F. Li, R. Zhang, Q. Li, S. Zhao, Res. Chem. Intermed. (2017). doi:10.1007/s11164-017-3001-9
9. S. Ganguly, P. Das, M. Bose, T.K. Das, S. Mondal, A.K. Das, N.C. Das, Ultrason. Sonochem. 39, 588
(2017)
10. V. Suvith, D. Philip, Spectrochim. Acta A 118, 532 (2014)
11. E. Akbarzadeh, M.R. Gholami, Res. Chem. Intermed. (2017). doi:10.1007/s11164-017-2965-9
12. M. Gopiraman, S. Saravanamoorthy, I.-M. Chung, Res. Chem. Intermed. (2017). doi:10.1007/
s11164-017-2950-3
13. D. Goodman, Catal. Lett. 99, 14 (2005)
14. A.A. Ashkarran, M. Ghavamipour, H. Hamidinezhad, H. Haddadi, Res. Chem. Intermed. 41, 7311
(2015)
15. X. Wang, B.G. Min, J. Hazard. Mater. 156, 386 (2008)
16. S.S. Metwally, I.M. Ahmed, H.E. Rizk, J. Alloys Compd. 709, 444 (2017)
17. M. Vallet-Regı́, J.M. González-Calbet, Prog. Solid State Chem. 32, 31 (2004)
18. S.V. Dorozhkin, Materials 2, 2045 (2009)
19. D.Y. Ji, T.F. Kuo, H.D. Wu, J.C. Yang, S.Y. Lee, Carbohydr. Polym. 89, 1130 (2012)
20. K. Lin, P. Liu, L. Wei, Z. Zou, W. Zhang, Y. Qian, Y. Shen, J. Chang, Chem. Eng. J. 222, 59 (2013)
21. C.J. Morrison, P. Gagnon, S.M. Cramer, Biotechnol. Bioeng. 108, 821 (2011)
22. M.R. Rogel, H. Qiu, G.A. Ameer, J. Mater. Chem. 18, 4241 (2008)
23. M. Wang, Biomaterials 24, 2151 (2003)
24. Z. Fang, Q. Feng, Mater. Sci. Eng. C 35, 194 (2014)
25. X. Wang, G. Song, T. Lou, Med. Eng. Phys. 32, 397 (2010)
26. F. Bernsmann, V. Ball, F. Addiego, A. Ponche, M. Michel, J.J.D.A. Gracio, V. Toniazzo, D. Ruch,
Langmuir 27, 2825 (2011)
27. H. Lee, S.M. Dellatore, W.M. Miller, P.B. Messersmith, Science 318, 430 (2007)
28. X. Cui, X. Chen, S. Chen, F. Jia, S. Yang, Z. Lin, Z. Shi, H. Deng, J. Alloys Compd. 693, 963 (2017)
29. L. Zhang, S. Yuan, S. Chen, D. Wang, B.-Z. Han, Z.-M. Dang, Compos. Sci. Technol. 110, 131
(2015)
30. H. Hu, B. Yu, Q. Ye, Y. Gu, F. Zhou, Carbon 48, 2353 (2010)
31. R. Sa, Z. Wei, Y. Yan, L. Wang, W. Wang, L. Zhang, N. Ning, M. Tian, Compos. Sci. Technol. 113,
62 (2015)
32. Y. Lee, T.G. Park, Langmuir 27, 2971 (2011)
33. M. Scampicchio, J. Wang, A.J. Blasco, A. Sanchez Arribas, S. Mannino, A. Escarpa, Anal. Chem. 78,
2063 (2006)
34. A. Ghorbani-Choghamarani, H. Rabiei, B. Tahmasbi, B. Ghasemi, F. Mardi, Res. Chem. Intermed.
42, 5737 (2016)
35. K.C. Black, Z. Liu, P.B. Messersmith, Chem. Mater. 23, 1135 (2011)
36. P. Wang, C. Li, H. Gong, X. Jiang, H. Wang, K. Li, Powder Technol. 203, 321 (2010)
37. J. He, X. Yang, J. Mao, F. Xu, Q. Cai, Appl. Surf. Sci. 258, 6830 (2012)
38. T.R. Mandlimath, B. Gopal, J. Mol. Catal. A: Chem. 350, 15 (2011)
39. J.P. Andreassen, J. Cryst. Growth 274, 264 (2005)
40. J. Yao, W. Tjandra, Y.Z. Chen, K.C. Tam, J. Ma, B. Soh, J. Mater. Chem. 13, 3057 (2003)
41. M.C. Chang, J. Tanaka, Biomaterials 23, 4858 (2002)
42. B. Fowler, E. Moreno, W. Brown, Arch. Oral Biol. 11, 492 (1966)
43. C. Durucan, P.W. Brown, J. Biomed. Mater. Res. 51, 725 (2000)
44. Y. Sun, Y. Deng, Z. Ye, S. Liang, Z. Tang, S. Wei, Colloids Surf. B 111, 116 (2013)
45. H. Sun, M. Ai, S. Zhu, X. Jia, Q. Cai, X. Yang, RSC Adv. 5, 95642 (2015)
46. I. Mobasherpour, M.S. Heshajin, A. Kazemzadeh, M. Zakeri, J. Alloys Compd. 430, 333 (2007)

123
T. K. Das et al.

47. I.S. Neira, Y.V. Kolen’ko, O.I. Lebedev, G. Van Tendeloo, H.S. Gupta, F. Guitián, M. Yoshimura,
Cryst. Growth Des. 9, 474 (2008)
48. S. Danilchenko, O. Kukharenko, C. Moseke, I.Y. Protsenko, L. Sukhodub, B. Sulkio-Cleff, Cryst.
Res. Technol. 37, 1240 (2002)
49. S. Jana, A.V. Kondakova, S.N. Shevchenko, E.V. Sheval, K.A. Gonchar, V.Y. Timoshenko, A.N.
Vasiliev, Colloids Surf. B 151, 254 (2017)
50. R.F. Elsupikhe, K. Shameli, M.B. Ahmad, Res. Chem. Intermed. 41, 8525 (2015)
51. M. Hao, M. Tang, W. Wang, M. Tian, L. Zhang, Y. Lu, Compos. B Eng. 95, 403 (2016)
52. Z. Yang, Y. Wu, J. Wang, B. Cao, C.Y. Tang, Environ. Sci. Technol. 50, 9550 (2016)
53. T. Zeng, H.Y. Niu, Y.R. Ma, W.H. Li, Y.Q. Cai, Appl. Catal. B 134, 33 (2013)
54. X. Huang, Y. Xiao, W. Zhang, M. Lang, Appl. Surf. Sci. 258, 2660 (2012)
55. N. Gupta, H.P. Singh, R.K. Sharma, J. Mol. Catal. A 335, 252 (2011)
56. M. Guo, J. He, Y. Li, S. Ma, X. Sun, J. Hazard. Mater. 310, 97 (2016)
57. K. Jiang, H.X. Zhang, Y.Y. Yang, R. Mothes, H. Lang, W.B. Cai, Chem. Commun. 47, 11926 (2011)
58. S. Jana, S.K. Ghosh, S. Nath, S. Pande, S. Praharaj, S. Panigrahi, S. Basu, T. Endo, T. Pal, Appl.
Catal. A 313, 48 (2006)
59. Y. Chi, Q. Yuan, Y. Li, J. Tu, L. Zhao, N. Li, X. Li, J. Colloid Interface Sci. 383, 102 (2012)
60. H. Liu, Q. Yang, J. Mater. Chem. 21, 32 (2011)
61. X.Q. Wu, D.D. Huang, Z.H. Zhou, W.W. Dong, Y.P. Wu, J. Zhao, D.S. Li, Q. Zhang, X. Bu, Dalton
Trans. 46, 2438 (2017)
62. L. Sooväli, E.I. Rõõm, A. Kütt, I. Kaljurand, I. Leito, Accredit. Qual. Assur. 11, 255 (2006)
63. F. Clare, Accredit. Qual. Assur. 10, 288 (2005)
64. S. Liu, Y. Cai, X. Cai, H. Li, F. Zhang, Q. Mu, Y. Liu, Y. Wang, Appl. Catal. A 453, 53 (2013)
65. B.R. Ganapuram, M. Alle, R. Dadigala, A. Dasari, V. Maragoni, V. Guttena, Int. Nano Lett. 5, 222
(2015)
66. W. Wang, F. Wang, Y. Kang, A. Wang, Chem. Eng. J. 237, 343 (2014)
67. L. Ćurković, D. Ljubas, H. Juretić, React. Kinet. Mech. Catal. 99, 208 (2010)
68. D. Kamel, A. Sihem, C. Halima, S. Tahar, Desalination 247, 422 (2009)
69. M. Graetzel, A.J. Frank, J. Phys. Chem. (US) 86, 2974 (1982)

123

You might also like