This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Paper The following article is Open access

Synthesis of nanoparticles composed of silver and silver chloride for a plasmonic photocatalyst using an extract from a weed Solidago altissima (goldenrod)

, , , , , and

Published 8 January 2016 © 2016 Vietnam Academy of Science & Technology
, , Citation Vemu Anil Kumar et al 2016 Adv. Nat. Sci: Nanosci. Nanotechnol. 7 015002 DOI 10.1088/2043-6262/7/1/015002

2043-6262/7/1/015002

Abstract

Phytosynthesis of nanomaterials is advantageous since it is economical, ecofriendly, and simple, and, what is more, in the synthetic protocols, nontoxic chemicals and biocompatible materials are used. Here, a green synthetic methodology of nanoparticles (NPs) composed of silver (Ag) and silver chloride (AgCl) NPs is developed using a leaf extract of Solidago altissima as a reducing agent for the first time. Utilization of a terrestrial weed for the synthesis of Ag and AgCl NPs is a novel environmentally friendly approach considering that no toxic chemicals, external halide source, or elaborate experimental procedures are included in the process. The optical properties and elemental compositions of as-synthesized Ag and AgCl NPs are well characterized, and the degradation of an organic dye, i.e., rhodamine B (RhB), is investigated using the Ag and AgCl NPs. We find that degradation of RhB is effectively achieved thanks to both surface plasmon resonance and semiconductor properties of Ag and AgCl NPs. The surface-enhanced Raman scattering and antibacterial activities are also examined. The present approach to the synthesis of NPs using a weed may encourage the utilization of hazardous plants for the creation of novel nanomaterials.

Export citation and abstract BibTeX RIS

Original content from this work may be used under the terms of the Creative Commons Attribution 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

It is a great challenge to develop a high performance photocatalyst, considering its application in the field of environmental technology, such as efficient degradation of organic pollutants [1]. Titanium dioxide (TiO2), which is a well-studied semiconductor, acts as an excellent photocatalyst in various areas of photodegradation, including self-cleaning and photoinduced superhydrophilicity [2, 3]. However, as its band gap is 3.0 eV for anatase and 3.2 eV for rutile, it requires irradiation of light of an ultraviolet (UV) range of wavelengths for the excitation of TiO2, which may lead to insufficient utilization of the solar spectra [4], excluding both visible (vis) and near infrared (NIR) regions [5]. Therefore, there is a need to synthesize a photocatalyst that can utilize vis light in the solar spectra and normal indoor light sources. Nanoparticles (NPs) composed of silver (Ag) and silver chloride (AgCl) NPs are acknowledged to be an efficient and stable photocatalyst under irradiation of vis light [610]. Ag NPs, which are one of the noble metallic ones, have a high scattering and enhancement of vis light due to the surface plasmon resonance (SPR) effect [11], whereas, AgCl NPs, which possess excellent electronic, magnetic, optical, and catalytic properties [12], act as a photosensitive semiconductor with a direct band gap of 5.15 eV (241 nm) and an indirect one of 3.25 eV (382 nm) [13, 14]. Ag and AgCl NPs would, therefore, act as an efficient plasmonic vis light photocatalyst [15, 16]. Photocatalysts are, in general, synthesized by the hydrothermal method [17], the in situ anion exchange method [18], and the in situ photoactive method [19]. However, some toxic chemicals, external halide sources, and additional stabilizing agents are required in the synthetic protocols, and the reaction time is rather long. Therefore, it is still necessary to invent some efficient ecofriendly method for synthesizing photocatalytic NPs.

Phytosynthesis of nanomaterials would be advantageous in such a sense that it is economical, ecofriendly, and simple, and, what is more, in the synthetic protocols, nontoxic chemicals and biocompatible materials are used [2022]. Synthesis of metal NPs using such green protocols is in great demand since no toxic chemicals are emitted into the environment [23, 24] and the synthetic devices can be scaled up easily since the procedures do not require any extreme conditions, such as high pressure, energy, and temperature. A variety of plant species has been used in the synthesis of NPs [25], and it is known that Ag and AgCl NPs can be phytosynthesized [26].

Weed species would, in general, cause hazardous situations on agricultural lands, which may eventually suppress native biodiversity and may trigger serious economic damage [27, 28]. A terrestrial weed; Solidago altissima (S. altissima), belonging to the Asteraceae (Compositae) family, is commonly called goldenrod. S. altissima is a native plant of North America and became a common alien plant in Japan several hundred years ago [29]. It is well known that S. altissima can grow in agricultural fields under various conditions, and the growth of rice seedlings, for example, can seriously be inhibited by S. altissima, which is caused mainly by allelochemicals, such as 2-cis-dehydromatricaria ester [30] and methyl-10-(2-methyl-2-butenoyloxy)-2-cis-8-cis-decadiene -4,6-diynoate [31], released from its rhizomes.

Here, we synthesize Ag and AgCl NPs based on a green protocol using an extract of S. altissima for the first time. The structures and elements of the particles are well characterized, and the photocatalytic activity of as-synthesized Ag and AgCl NPs is examined using rhodamine B (RhB). We find that the degradation of RhB is achieved effectively thanks to both SPR and semiconductor properties of Ag and AgCl NPs. The surface-enhanced Raman scattering (SERS) and antibacterial activities are also investigated. We suppose that the present approach to the synthesis of NPs using a weed may encourage the utilization of hazardous plants for the creation of novel nanomaterials and that this may help in the battle to conserve nature.

2. Experimental section

2.1. Phytosynthesis of Ag and AgCl NPs

Fresh plant leaves of S. altissima were collected on the Kawagoe Campus, Toyo University, Japan and were cleaned with millipore water to remove the surface contaminants. Some 10 g of leaf pieces were heated in 100 ml of millipore water at 90 °C for 90 min, followed by filtration (Whatman Cat No 1002-185) to obtain an aqueous plant leaf extract. The leaf extract was refrigerated at 4 °C for further use. Some 20 ml of the aqueous leaf extract was mixed with 60 ml of 1 mM aqueous solution of Ag nitrate (AgNO3) (Sigma-Aldrich, Japan) at 4 °C, and the mixture was stirred for 30 min in the dark and then was exposed to light (power density: 1 sun = 100 mW cm−2, exposed area: 3 × 3 cm2) (HAL 320, Asahi spectra Co., Ltd.) for 30 min. The mixture was centrifuged at 15 000 rpm for 40 min to pelletize the product. The pellet was purified by dispersing it in millipore water, followed by centrifugation at 15 000 rpm for 40 min.

2.2. Characterization of Ag and AgCl NPs

The optical properties of as-synthesized NPs were characterized by an UV–vis–NIR spectrophotometer (DU 730, Beckman Coulter) in the range from 200 to 1000 nm at a resolution of 1 nm and a UV diffuse reflectance spectroscope (DH2000-DUV, USB2000, Ocean Optics, Inc.) in the range from 250 to 600 nm (figure 1). The morphology of the particles was analyzed using a transmission electron microscope (TEM) (JEM2200FS, JEOL) at an acceleration voltage of 200 kV with a scan snap, orius camera, and software (Gatan). Samples for TEM observation were prepared on a carbon (C) coated copper (Cu) grid (Cu 200 mesh, JEOL), dropping and drying NP suspension on the grid. The elemental composition of the particles was analyzed by energy dispersive x-ray spectroscopy (EDS) (SU6600, Hitachi equipped with OXFORD X-MaxN) (figure 2). The crystal nature of the particles was determined by an x-ray diffractometer (XRD) (Smart Laboratory, Rigaku), using a Cu x-ray of a 0.154 nm wavelength supplied by a 9 kW rotating Cu anode x-ray generator and employing a grazing-incidence x-ray diffraction geometry with a scanning diffraction angle (2θ) from 25° to 90° at 1.5° min−1 at an incident angle of 0.1°. The elemental and chemical bonding states of the NPs were studied by an x-ray photoelectron spectroscope (XPS) (Axis-HIs, Kratos Analytical) equipped with a nonmonochromatic Al x-ray of an energy of 1486.61 eV, pass energy of 80 eV, and step energy of 0.1 eV at lower than 1.0 × 10−8 torr (figure 3). Samples for XPS were prepared on a hydrofluoric acid treated silicon (Si) wafer, dropping and drying NP suspension on the wafer. The photoluminescence (PL) spectra of the particles were determined using a spectrofluorometer (FP 6500, Jasco).

Figure 1.

Figure 1. UV–vis–NIR spectra of a reaction mixture at regular intervals of time. The absorption peak at 462 nm increased gradually, representing the growth.

Standard image High-resolution image
Figure 2.

Figure 2. SEM, EDS, TEM, and FFT images of Ag and AgCl NPs. (a) SEM image of clusters of nanoparticle, (b) EDS spectrum of Ag and AgCl NPs, and the inset showing atomic and weight percentages of Ag and Cl. Si originates from the substrate, (c) TEM image of a cluster, and (d) TEM image of the region represented by the red square in image (c). The gaps between two neighboring fringes are 0.23 and 0.27 nm, which coincide with those of Ag (111) and AgCl (200), respectively; (e), (f) FFT images of the regions represented by the yellow and red squares in image (d). It is clearly shown that each cluster is composed of Ag and AgCl.

Standard image High-resolution image
Figure 3.

Figure 3. XRD of as-synthesized nanoparticles. The data also confirm that the clusters are composed of Ag and AgCl.

Standard image High-resolution image

2.3. Characterization of a leaf extract

The chemical functional groups of the organic compounds responsible for the reduction reaction were analyzed by Fourier transform infrared spectroscopy (FTIR) (Nicolet iS50, Thermo Fischer Scientific), scanning the IR from 400 to 4000 cm−1 at a resolution of 4 cm−1. The leaf extract was centrifuged at 15 000 rpm for 45 min, the supernatant was removed, and the pellet was dried in vacuum. Samples for the FTIR measurement were prepared by the KBr pellet method [32]. The elemental composition of the leaf extract was analyzed by EDS and XPS. Samples for EDS and XPS were prepared on a hydrofluoric acid treated Si wafer, dropping and drying an aqueous leaf extract on the wafer.

2.4. Measurement of the photocatalytic activity

Some 2 mg of NPs were mixed with 2 ml of a 10 mg l−1 aqueous solution of RhB (Tokyo Chemical Industry Co., Ltd.) for 30 min in the dark for the establishment of adsorption/desorption equilibrium of RhB on the particles. Then, the mixture was transferred to a quartz photoreactor for the measurement of the photocatalytic activity of the nanoparticles. The photoreactor was irradiated with light from a solar simulator without any cutoff filters (power density: 1 sun = 100 mW cm−2 and an exposed area of 3 × 3 cm2) (HAL 320, Asahi Spectra Co., Ltd.) ranging from 350 to 1100 nm to evaluate the degradation of RhB. Aliquots of the reaction mixture were collected and were centrifuged at 15 000 rpm for 10 min to separate the photocatalyst from the mixture, and UV–vis–NIR absorption by the supernatant was recorded by the UV–vis–NIR spectrophotometer for 60 min at regular intervals of time to measure the degradation of RhB (figure 4). Degraded RhB was further confirmed by a mass spectroscope (amaZon speed, Brüker Daltonics) by the direct probe method using an atmospheric pressure chemical ionization source at the mass-to-charge ratio (m/z) ranging from 15 to 600 amu. For examining the enhancement of the photocatalytic efficiency of the present NPs, another experiment was carried out using a commercialized photocatalyst Ag3PO4, the size of which varied from 25 to 600 nm (Sigma-Aldrich, Japan) under the same experimental conditions.

Figure 4.

Figure 4. Degradation of RhB. (a) UV–vis–NIR absorption spectrum of RhB degraded by Ag and AgCl NPs at regular intervals of time, (b) UV–vis–NIR absorption spectrum of RhB degraded by Ag phosphate (Ag3PO4), and (c) time variation of —ln (C/C0), where C is the concentration of the dye at time t and C0 is the concentration at t = 0.

Standard image High-resolution image

2.5. SERS analysis

The SERS characteristics were investigated using an aqueous solution of RhB. A 0.1 mM RhB solution was dropped onto dried Ag and AgCl NPs placed on a glass substrate, and the Raman spectra were obtained using a micro-Raman spectroscopic system (Laboratory Ram HR800UV, Horiba Jobin Yvon S.A.S.) with an excitation wavelength of 633 nm (figure 5). The spot diameter and power of the laser beam were 2 μm and 43.6 μW, respectively, and the spectra were obtained with a 5 s integration.

Figure 5.

Figure 5. Raman spectra of a 0.1 mM aqueous solution of RhB. The red curve represents a Raman spectrum of RhB solution dropped on dried Ag and AgCl NPs, whereas, the blue curve represents the spectrum of the solution dropped on a glass substrate. The Raman spectrum of RhB on Ag and AgCl NPs shows several sharp peaks, which are typical ones of RhB molecules.

Standard image High-resolution image

2.6. Antibacterial activity

To examine the antibacterial activity of the Ag and AgCl NPs, the time variation of the bacterial growth was measured for 15 h, culturing two bacterial species; Escherichia coli (gram-negative) and Bacillus subtilis (gram-positive) in nutrient broth No. 2 (Oxoid) media supplemented with an aqueous solution of Ag and AgCl NPs, the concentration of which was changed, 5, 10, and 20 μg ml−1 at 37 °C under a constant shaking at 150 rpm. The optical density (O.D.) values at 600 nm were recorded every 30 min using an O.D. monitor (C&T, Taitec Corp.).

3. Results and discussion

3.1. Synthesis of Ag and AgCl NPs

We successfully synthesized Ag and AgCl NPs by the following two-step procedure; (1) the reduction of Ag+ (AgNO3) using an aqueous leaf extract of S. altissima to synthesize AgCl NPs and (2) the photoreduction of AgCl NPs to synthesize Ag and AgCl NPs. The chemical compounds present in the aqueous leaf extract, such as terpenoids, glycosides, acetylenes, phenols, and components containing Cl [29, 3338] were supposed to be responsible for the reduction reaction to synthesize AgCl NPs in the first procedure [39, 40] noting that no external Cl components or any additional chemicals were supplied during the synthesis (see figure A1 in the appendix for FTIR, EDS, and XPS analyses of the leaf extract). The second procedure was performed by irradiation of light into the reaction mixture. Ag and AgCl NPs were synthesized owing to the high photosensitivity of AgCl [6, 41]. A pair of electrons and holes is generated in AgCl via photon activation, and, therefore, it is supposed that the photogenerated electron promoted the conversion of the Ag+ ion into the Ag0 atom [7, 41, 42], resulting in the synthesis of Ag and AgCl NPs. The progress of the synthesis of Ag and AgCl NPs was confirmed by the time variation of the UV–vis–NIR spectrum of the reaction mixture during the synthetic procedure. As shown in figure 1, the time variation of the peak intensity observed in the vis region depicts the gradual synthesis of Ag and AgCl NPs. Figure A2 in the appendix shows the UV–vis–NIR spectrum of the purified Ag–AgCl NPs dispersed in millipore water where the plasmonic resonance peak at 462 nm is attributed to Ag, whereas, the absorption peak at 246 nm is attributed to AgCl [43]. The UV–vis diffuse reflectance spectrum of dried Ag and AgCl NPs shows some absorbance in the vis light region due to SPR in metallic Ag clusters in the powders [44], while the absorbance in the UV region is caused by AgCl NPs coupled with AgNPs (see figure A3 in the appendix) [45], which suggests that the present Ag and AgCl NPs can be used as an effective plasmonic photocatalyst in vis light [46]. The PL spectrum of the Ag and AgCl NPs at room temperature (24 ± 2 °C) showed emission bands at 318 and 470 nm when excited with photons of a 200 nm wavelength, which also clarifies the semiconductor features of the present Ag and AgCl NPs (see figure A4 in the appendix).

The morphology and crystalline nature of the particles were investigated by SEM, EDS, TEM, and XRD. SEM, EDS, TEM, and FFT images of the particles are shown in figure 2. A large number of clusters of particles, which were composed of Ag and Cl, were produced (see figures 2(a) and (b)). It is clearly shown that each cluster was composed of Ag and AgCl (figures 2(c)–(f)), noting that the lattice fringe distance of 0.23 nm corresponds to Ag (111), whereas, 0.27 nm corresponds to AgCl (200) (figures 2(d)–(f)). The peaks in the diffractogram also confirm the cubic AgCl and cubic Ag structures, respectively, with the lattice constants of 5.553 050 Å (ICDD 01-071-5209) and 4.097 667 Å (ICDD 00-041-1402) (see figure 3).

The elemental composition of the as-synthesized NPs was analyzed by XPS (see figure A5 in the appendix). The wide scan spectrum indicates the presence of Ag, Cl, C, and O (see figure A5(a)). Ag3d5/2 and Ag3d3/2 centred at 367.75 and 373.76 eV represent AgCl (figure A5(b)) [47, 48], whereas, Cl2p3/2 and Cl2p1/2 centred at 197.90 and 199.57 eV represent Cl present in Ag and AgCl NPs (figure A5(c)) [49, 50].

3.2. Photocatalytic performance of Ag and AgCl NPs

The photocatalytic efficiency of the as-synthesized Ag and AgCl NPs was evaluated via the degradation of an organic dye RhB (10 mg l−1) under the irradiation of light generated from the solar simulator. RhB molecules in an aqueous solution show three major peaks at 258, 354, and 554 nm in the UV–vis absorption spectrum. Figures 4(a) and (b) show the time variations of the absorption spectra in the presence of Ag and AgClNPs, and Ag3PO4NPs. The decrease in the peak intensity and the shift of the peak position toward shorter wavelength regions elucidate the degradation of RhB molecules and N-deethylation [51]. The photocatalytic activity was evaluated by the UV–vis absorption at 554 nm. The time variation of −ln (C/C0), where C is the concentration of the dye at regular intervals of time t and C0 is the concentration at t = 0, is shown in figure 4(c). The rate constant k for the degradation reaction of RhB based on the first-order reaction kinetics (ln (C/C0) = −kt) is calculated to be 0.0387 min−1 in the case of Ag and AgCl NPs, whereas, k = 0.0237 min−1 in the case of Ag3PO4. The high photocatalytic performance of Ag and AgCl NPs owes to the SPR of AgNPs and the semiconductor features of AgCl NPs [47, 52]. A further confirmation of the degradation of RhB by Ag and AgCl NPs was carried out by mass spectrometric analysis. Figure A6 in the appendix shows the peaks at 442, 413, 386, 357, and 329 amu, which correspond to RhB and its N-deethylated intermediates [53]. It is clearly shown that the degradation of RhB is encouraged by the present Ag and AgCl NPs.

The change in the ratio of Ag and AgCl in the Ag and AgCl NPs before and after the photocatalytic experiment was evaluated by XRD analysis (see figure A7 in the appendix). The peak intensities in the XRD spectrum and weight percentages of Ag and AgCl changed quite significantly after the experiment due to the photosensitivity of AgCl, noting that no cutoff filter was used during light irradiation in the present study. The previous studies showed that Ag NPs on the AgCl NPs were structurally stable and catalytically reusable thanks to the cutoff filter of UV light [7, 54], and therefore, we believe that the catalytic performance of the present Ag and AgCl will be improved by introducing the cutoff filter of UV light.

3.3. SERS activity of Ag and AgCl NPs

Raman spectra of an aqueous solution of RhB, which was dropped onto Ag and AgCl NPs placed on a glass substrate, was obtained (see figure 5). It is clearly shown that Raman signals were greatly enhanced in the presence of Ag and AgCl NPs, noting that each peak coincides with that of RhB [55] and, therefore, the present Ag and AgCl NPs are highly SERS active thanks to Ag NPs [56, 57], whereas, there were no recognizable signals of RhB when the solution was dropped onto the surface of a glass substrate.

3.4. Antibacterial activity

The antibacterial activity of the Ag and AgCl NPs was investigated culturing Escherichia coli (gram-negative) and Bacillus subtilis (gram-positive) in the presence of Ag and AgCl NPs in the liquid medium. The time variation of the bacterial growth for 15 h is shown in figure A8 in the appendix. The growth of Escherichia coli and Bacillus subtilis in the presence of Ag and AgCl NPs was reduced even when the concentration of Ag and AgCl NPs was as low as 5 μg ml−1 [58, 59].

4. Conclusions

We developed a green method for synthesizing an effective plasmonic photocatalyst based on a two-step procedure, i.e., (1) production of AgCl NPs via the reduction of Ag+ (AgNO3) using an aqueous leaf extract of S. altissima, and (2) production of Ag and AgCl NPs via the photoreduction of AgCl NPs. The optical properties and elemental compositions of the as-synthesized Ag and AgCl NPs were well characterized. We found that degradation of RhB was effectively achieved thanks to both SPR and semiconductor properties of Ag and AgCl NPs. The particles also showed high surface-enhanced Raman scattering and antibacterial activities. The present green approach to the synthesis of NPs using a weed may encourage the utilization of hazardous plants for the creation of novel nanomaterials.

Acknowledgments

Part of the present study was supported by a Grant for the Programme for the Strategic Research Foundation at Private Universities, S1101017, organized by the Ministry of Education, Culture, Sports, Science and Technology (MEXT), Japan since April 2011. We would like to thank K Yanagisawa, K Hirakawa, H Sato, Bio-Nano Electronics Research Centre, Toyo University for their support for the TEM observation, and Dr N Hirotsu, Faculty of Life Sciences, Toyo University for his support for designing the experiment using S. altissima. V A K would like to thank MEXT for their financial support since April 2013.

Appendix

A1.: FTIR, EDS, and XPS spectra of the leaf extract

FTIR spectrum of a leaf extract is shown in figure A1(a). The peak position at 3354.07 cm−1 corresponds to a N–H stretch, the peak position at 2921.79 and 2851.17 cm−1 corresponds to a C–H stretch, the peak position at 1640.75 cm−1 corresponds to a N–H bend, the peak position at 1384.19 and 1327.46 cm−1 corresponds to S=O, the peak position at 1156.92 and 1076.78 cm−1 corresponds to C–O, the peak position at 781.52 cm−1 corresponds to C–Cl, and the peak position at 611.77 and 521.64 cm−1 corresponds to C–X.1 Note that the peak at 781.52 cm−1 represents Cl related compounds. The EDS spectrum of a leaf extract is shown in figure A1(b). The peak positions represent P, sulfur, and Cl that were present in the leaf extract. The XPS data, i.e., the peaks of C1s, O1s, and Cl2p, show that the present leaf extract was composed of some compounds bonded with C, O, and Cl (see figure A1(c)). Figure A1(d) shows two peaks corresponding to Cl2p3/2 and Cl2p1/2 centered at 198.29 and 199.95, which represent Cl related components are present in the leaf extract.2 The peaks of C1s correspond to binding energies of 284.6, 286.06, 293.09, and 295.72 eV (see figure A1(e)).

Figure A1.

Figure A1. Characterization of a leaf extract. (a) FTIR spectrum, (b) EDS spectrum, and (c)–(e) XPS spectra.

Standard image High-resolution image

A2.: UV–vis–NIR spectrum of purified Ag and AgCl NPs

Figure A2.

Figure A2. Absorption spectrum of purified Ag and AgCl NPs dispersed in millipore water.

Standard image High-resolution image

A3.: UV–vis diffuse reflectance spectrum of dried Ag and AgCl NPs

Figure A3.

Figure A3. Absorption spectrum of dried Ag and AgCl NPs on a Si substrate.

Standard image High-resolution image

A4.: PL spectrum of Ag and AgCl NPs

Figure A4.

Figure A4. PL spectrum of Ag and AgCl NPs. The emission peaks at 318 and 470 nm clarify the semiconductor property of Ag and AgCl. The peaks indicated by asterisks are caused by higher-order diffraction.

Standard image High-resolution image

A5.: XPS spectra of Ag and AgCl NPs

Figure A5.

Figure A5. XPS spectra of as-synthesized particles. (a) Wide scan spectrum. The peaks indicated by asterisks represent unassigned Auger ones, (b) Ag3d, and (c) Cl2p.

Standard image High-resolution image

A6.: Mass spectra of degraded RhB

Figure A6.

Figure A6. Mass spectra of RhB degraded by Ag and AgCl NPs at regular time intervals.

Standard image High-resolution image

A7.: XRD analysis of Ag and AgCl NPs

Figure A7.

Figure A7. XRD analysis of Ag and AgCl NPs before and after the photocatalysis experiment. (a) XRD of Ag and AgCl NPs before (blue) and after the experiment (red), (b) intensities at different peak positions, representing various crystal lattice planes of a  diffractogram of Ag and AgCl before (blue) and after the experiment (red), and (c) weight percentages of AgCl and Ag before (blue) and after the experiment (red).

Standard image High-resolution image

A8.: Growth of bacteria in the presence of Ag and AgCl NPs

Figure A8.

Figure A8. Time variations of the growth of (a) Escherichia coli and (b) Bacillus subtilis in the presence of Ag and AgCl NPs.

Standard image High-resolution image
Please wait… references are loading.