One-pot green synthesis of silver nanocrystals using Hymenodictyon orixense: a cheap and effective tool against malaria, chikungunya and Japanese encephalitis mosquito vectors?

Marimuthu Govindarajan*a and Giovanni Benelli*b
aUnit of Vector Control, Phytochemistry and Nanotechnology, Department of Zoology, Annamalai University, Annamalainagar-608 002, Tamil Nadu, India. E-mail: drgovind1979@gmail.com; Fax: +91 04144 238080; Tel: +91 9585265999
bInsect Behaviour Group, Department of Agriculture, Food and Environment, University of Pisa, via del Borghetto 80, 56124 Pisa, Italy. E-mail: g.benelli@sssup.it; benelli.giovanni@gmail.com; Fax: +39 0502216087; Tel: +39 0502216141

Received 20th April 2016 , Accepted 13th June 2016

First published on 15th June 2016


Abstract

Mosquitoes are important vectors of malaria, dengue, Zika virus and many other parasites and pathogens of public health relevance. Recently, the green nanosynthesis of mosquitocides relying on plant compounds as reducing and stabilizing agents has received growing interest, due to the absence of toxic chemicals and high-energy input. In this research, Hymenodictyon orixense-mediated synthesis of silver nanoparticles (AgNPs) was conducted to control larval populations of the malaria vector Anopheles subpictus, the chikungunya vector Aedes albopictus and the Japanese encephalitis vector Culex tritaeniorhynchus. AgNPs were characterized using UV-visible spectrophotometry, FTIR spectroscopy, EDX and XRD analyses, AFM, SEM and TEM. AgNPs were toxic towards all the mosquito vectors, LC50 values ranged from 17.10 μg ml−1 to 20.08 μg ml−1. Notably, AgNPs were safer to the non-target mosquito predator Diplonychus indicus (LC50 = 833 μg ml−1). Overall, H. orixense-fabricated AgNPs can be considered for the development of novel and safer control tools against mosquito vectors of medical and veterinary importance.


Introduction

Mosquitoes (Diptera: Culicidae) are responsible for the transmission of important and dreadful pathogens and parasites worldwide, including malaria, dengue, yellow fever, filariasis and Zika virus.1 Mosquito-borne diseases are endemic in over 100 countries, causing mortality of nearly two million people every year, and at least one million children die of such diseases each year, leaving as many as 2100 million people at risk around the world.2–4 Chemical insecticides are widely used to control mosquitoes. However, they are often harmful to non-target organisms and human health.5,6 Therefore, effective and eco-friendly control strategies are urgently needed. The development of plant-borne pesticides with multiple mechanisms of action may be successful for mosquito control.7 Current emphasis on bioactivity of plant materials highlighted excellent toxic properties against different developmental instars of a number of Culicidae vectors.8

Furthermore, with the progress of nanotechnology, the production of metal nanoparticles for mosquitocidal purposes has been studied.9 In recent years, the nanosynthetic methods based on the employment of plant extracts received more attention over chemical, physical and microbial-based ones, due to the absence toxic chemicals and high-energy inputs. Recently, a growing number of plants have been successfully used for efficient and rapid extracellular synthesis of silver, copper, and gold nanoparticles.10 Good examples include cheap extracts of neem, Azadirachta indica,11 Chomelia asiatica,12 Sida acuta,13 Gmelina asiatica,14 and Pongamia pinnata.15 Interestingly, nanoparticles possess peculiar toxicity mechanisms due to surface modification.16 Silver nanoparticles also have antibacterial, antifungal, antiplasmodial and mosquitocidal properties.17 However, despite the increasing number of evidences of plant-synthesized mosquitocidal nanoparticles, only moderate efforts have been carried out to shed light on the nanoparticle toxicity on non-target organisms sharing the same ecological niche of mosquito young instars.9

Hymenodictyon orixense (Roxb.) Mabb. (Rubiaceae) is native to tropical Asia and widely distributed in India. It is a tree commonly known as “Vellai kadambu” in Tamil Nadu region. Leaves are decussate, elliptic-ovate, and acute-attenuate at base. Flowers in terminal and axillary paniculate raceme, faintly scented, greenish. Capsules ellipsoid, on recurved pedicels. The stem bark contains tannin, toxic alkaloid, hymenodictine, a bitter substance, aesculin, an apioglucoside of scopoletin, hymexelsin.18 Anthraquinones, rubiadin and its methyl ether, lucidin, nordamnacanthal, damnacanthal, 2-benzylzanthopurpurin, anthragallol, soranjidol and morindone have been isolated from roots.19

In this research, we reported a method to synthesize silver nanoparticles (Ag NPs) using the aqueous leaf extract of the H. orixense, a cheap and eco-friendly material acting as reducing and stabilizing agent. Ag NPs were characterized by UV-visible spectrophotometry, Fourier transform infrared spectroscopy (FTIR), energy dispersive X-ray analysis (EDX), X-ray diffraction analysis (XRD), atomic force microscopy (AFM), scanning electron microscopy (SEM) and transmission electron microscopy (TEM). The aqueous extract of H. orixense and the synthesized Ag NPs were tested for their larvicidal potential against the malaria vector Anopheles subpictus, the chikungunya vector Aedes albopictus and the Japanese encephalitis vector Culex tritaeniorhynchus. Furthermore, we evaluated the biotoxicity of H. orixense aqueous extract and green-synthesized Ag NPs on the non-target aquatic predator Diplonychus indicus, which shares the same ecological niche of Anopheles and Aedes mosquitoes in India.

Experimental

Materials

Silver nitrate was purchased from Merck, India. The glassware was acid-washed thoroughly and then rinsed with Millipore Milli-Q water. Healthy and fresh leaves of H. orixense were collected from Nilgiris, Western Ghats (11° 10′ N to 11° 45′ N latitude and 76° 14′ E to 77° 2′ E longitude), Tamil Nadu state, India. The identity was confirmed at the Department of Botany, Annamalai University, Annamalai Nagar, Tamil Nadu. Voucher specimens were numbered and kept in our laboratory and are available under request (ID: AUDZ-318).

Preparation of plant leaf extract

The leaves of H. orixense were dried in the shade and ground to fine powder in an electric grinder. Aqueous extract was prepared by mixing 50 g of dried leaf powder with 500 ml of water (boiled and cooled distilled water) with constant stirring on a magnetic stirrer. The suspension of dried leaf powder in water was left for 3 h and filtered through Whatman no. 1 filter paper and the filtrate was stored in an amber-coloured airtight bottle at 10 °C temperature until testing.

Synthesis and bio-physical characterization of silver nanoparticles

The broth solution of fresh leaves was prepared by taking 10 g of thoroughly washed and finely cut leaves in a 300 ml Erlenmeyer flask along with 100 ml of sterilized double-distilled water and then boiling the mixture at 1000 °C for 5 min before finally decanting it for 30 min. The extract was filtered with Whatman filter paper no. 1, stored at −15 °C, in order to prevent degradation processes, and tested within 7 days. The filtrate was treated with aqueous 1 mM AgNO3 (21.2 mg of AgNO3 in 125 ml of Milli-Q water) solution in an Erlenmeyer flask and incubated at room temperature (28 ± 2 °C, R.H. 70–85%). Eighty-eight millilitres of an aqueous solution of 1 mM silver nitrate was reduced using 12 ml of leaf extract at room temperature for 10 min, resulting in a brown-yellow solution indicating the formation of Ag NPs.

The bioreduction of Ag+ ions was monitored using a UV-Vis spectrophotometer (UV-160v, Shimadzu, Japan). Analysis on size, morphology, agglomeration pattern and dispersed nature of Ag NPs were performed by atomic force microscopy (Agilent Technologies AFM-5500), scanning electron microscopy (Hitachi S3000 H SEM) and transmission electron microscopy (TEM Technite 10 Philips). The purified Ag NPs were examined for the presence of biomolecules using FTIR spectroscopy (Thermo Scientific Nicolet 380 FT-IR Spectrometer) KBr pellets and crystalline Ag NPs were determined by XRD analysis.

Mosquito rearing

Following the method by Govindarajan and Benelli20 laboratory-bred pathogen-free strains of mosquitoes were reared in the vector control laboratory, Department of Zoology, Annamalai University. At the time of adult feeding, these mosquitoes were 3–4 days old after emergences (maintained on raisins and water) and were starved for 12 h before feeding. Each time, 500 mosquitoes per cage were fed on blood using a feeding unit fitted with Parafilm as membrane for 4 h. A. albopictus feeding was done from 12 noon to 4.00 p.m. and A. subpictus and C. tritaeniorhynchus were fed during 6.00 p.m. to 10.00 p.m. A membrane feeder with the bottom end fitted with Parafilm was placed with 2.0 ml of the blood sample (obtained from a slaughterhouse by collecting in a heparinized vial and stored at 4 °C) and kept over a netted cage of mosquitoes. The blood was stirred continuously using an automated stirring device, and a constant temperature of 37 °C were maintained using a water jacket circulating system. After feeding, the fully engorged females were separated and maintained on raisins. Mosquitoes were held at 28 ± 2 °C, 70–85% relative humidity, with a photoperiod of 12 h light and 12 h dark.

Acute toxicity against mosquito larvae

Larvicidal activity of the aqueous extract and Ag NPs from H. orixense was evaluated according to WHO protocol.21 Based on the wide range and narrow range tests, aqueous crude extract was tested at 50, 100, 150, 200 and 250 μg ml−1 concentrations and Ag NP was tested at 8, 16, 24, 32 and 40 μg ml−1 concentrations. Twenty numbers of late III instar larvae were introduced into a 500 ml glass beaker containing 249 ml of dechlorinated water, and 1 ml of desired concentrations of leaf extract or Ag NP was added. For each concentration, five replicates were performed. Larval mortality was recorded at 24 h after exposure, during which no food was given to the larvae. Each test included a set control groups (i.e. silver nitrate tested at the same concentrations of green-fabricated Ag NP, and distilled water) with five replicates for each individual concentration.

Biosafety on non-target mosquito predators

The effect of non-target organisms was assessed following the method by Sivagnaname and Kalyanasundaram.22 The effect of aqueous extract and Ag NPs of the potential plant was tested against non-target organisms D. indicus. The species were field collected and separately maintained in cement tanks (85 cm diameter and 30 cm depth) containing water at 27 ± 3 °C, air relative humidity was 85%.

The aqueous extract and Ag NPs of H. orixense were evaluated at a concentration of even 50 times higher the LC50 dose for mosquito larvae. Ten replicates will be performed for each concentration along with four replicates of untreated controls. Non-target organisms were observed continuously for 10 days to understand the post-treatment effect of the extract and Ag NP on their survival.

Data analysis

Data were analysed using the SPSS Statistical Software Package version 16.0. For all tested species, toxicity data were subjected to probit analysis. LC50 and LC90 were calculated using the method by Finney.23 For each LC50 and LC90 calculated in acute toxicity assays, the 95% confidence limits, regression equation, slope and chi square values were provided. In experiments evaluating the acute toxicity on non-target organism and the Suitability Index (SI) was calculated for non-target species using the following formula.24
image file: c6ra10228j-t1.tif

Results and discussion

Synthesis and characterization of silver nanoparticles

The colour of H. orixense aqueous extract changed from yellowish to dark brown after the reduction of silver nitrate (Fig. 1a). The colour change indicates Ag+ reduction to elemental nanosilver. Fig. 1b shows the UV-visible spectrum of biosynthesized silver nanoparticles after 180 min from the reaction. An intense broad absorption peak was observed at 449 nm, and it was probably due to surface plasmon resonance (SPR). The SPR peak is very sensitive to the size and shape of the nanoparticles, amount of extract, silver nitrate concentration and the type of biomolecules present in the leaf extract (Fig. 1b). Our UV-Vis results are in agreement with previous research,10,25,26 where the Ag NPs were observed as stable in solution and also showed little aggregation. Besides, the plasmon bands were broadened with an absorption tail in longer wavelengths, and this may be related to the size distribution of nanoparticles.27
image file: c6ra10228j-f1.tif
Fig. 1 (a) Color intensity of Hymenodictyon orixense aqueous extract before and after the reduction of silver nitrate (1 mM). The color change indicates Ag+ reduction to elemental nanosilver. (b) UV-visible spectrum of silver nanoparticles after 180 min from the reaction.

The crystalline nature of Ag NPs was studied by XRD analysis (Fig. 2). The XRD pattern confirmed the crystalline nature of bio-fabricated Ag NP. Four diffraction peaks were observed at 38.22, 44.37, 64.54 and 77.47 represent the (111), (200), (220) and (311), reflections and the face-centered cubic structure of metallic silver, respectively.9 Notably, limited information is available on the phytoconstituents of the extract of H. orixense leaves. Probably, hymenodictine, aesculin, scopoletin and hymexelsin are major compounds, as already reported for the bark of this species.18,28 The FTIR spectrum of synthesized silver nanoparticles by using H. orixense leaf extract is shown in Fig. 3. It is confirmed the fact that to identify the biomolecules for reduction and efficient stabilization of the metal nanoparticles. The band at 3420.95 cm−1 may correspond to O–H, as also the H-bonded alcohols and phenols. Shanmugam et al.29 suggested that these bonds could be due to the stretching of –OH in proteins, enzymes or polysaccharides present in the extract. The peak at 2922 cm−1 may indicate the presence of carboxylic acids.30 2851 cm−1 may be linked to the presence of methylene groups (CH2) close to OH or NH2 functional groups. Similarly, bands at 1653 cm−1, 1636 cm−1 and 1457 cm−1 could be due to the C‚C stretching of aromatic rings, which may indicate the presence of benzene ring, a band at very low intensity at 1684 cm−1 suggests the ortho-substitution pattern of benzene ring. In addition, shoulder peaks at 1559 cm−1, 1541 cm−1, and 1507 cm−1 probably indicate that the amide I and amide II probably arise due to carbonyl and –NH stretch vibrations in the amide linkages of the proteins, respectively. The band at 1384 cm−1 may correspond to C–C stretching of aromatic amines. The bands at 1111.07 cm−1, and 1019.16 cm−1 might indicate the presence of C–O stretching alcohols, carboxylic acids, esters and ethers, while 693.43 cm−1 probably corresponds to C–H stretching strong vinyl di-substituted alkenes.31


image file: c6ra10228j-f2.tif
Fig. 2 XRD pattern of silver nanoparticles synthesized using the Hymenodictyon orixense aqueous extract.

image file: c6ra10228j-f3.tif
Fig. 3 FTIR spectrum of silver nanoparticles synthesized using the Hymenodictyon orixense aqueous leaf extract.

AFM is a primary tool for analysing size, shape, agglomeration pattern and offers visualizations of three-dimensional views of the nanoparticles unlike the electron microscopes. It has an advantage over combination of high resolution, samples does not have to be conductive and does not require the high-pressure vacuum conditions. Our 2.5 μm resolution studies of biofabricated synthesized Ag NPs with AFM revealed the particles are poly-dispersed, spherical in shape, with size ranging from 0.3 to 3.5 nm. No agglomeration was observed between the particles (Fig. 4a). Raw data obtained from this AFM microscope were treated with a specially designed image processing software (NOVA-TX) to further exploit the 3D image of nanoparticles (Fig. 4b). The average nanoparticle size obtained from the corresponding diameter distribution was about 2 nm (Fig. 4c and d). SEM of Ag NPs showed that the nanoparticles were mostly spherical (Fig. 5a). Fig. 5b showed the TEM of Ag NP synthesized using H. orixense leaf extract, highlighting that some nanoparticles were of a size slightly bigger (i.e. 25–35 nm) to the mean size reported in AFM size analysis, however, still below 100 nm. We also noted that “capped” Ag NP were stable in solution for at least 8 weeks. In addition, EDX analysis provided information on the chemical analysis at specific locations of the sample (spot EDX). Fig. 6 shows spot EDX analysis, confirming the presence of silver in the sample.


image file: c6ra10228j-f4.tif
Fig. 4 Atomic force microscopy (AFM) of silver nanoparticles green synthesized from Hymenodictyon orixense, (a) 2.5 μm resolution studies of 0.3 to 3.5 nm size, spherical shaped, poly-dispersed particles, (b) 3D image of silver nanoparticles analyzed by NOVA-TX software, (c) histogram and (d) line graph showing the size distribution of green-synthesized silver nanoparticles.

image file: c6ra10228j-f5.tif
Fig. 5 (a) Scanning electron micrograph and (b) transmission electron micrograph of the Hymenodictyon orixense-synthesized silver nanoparticles.

image file: c6ra10228j-f6.tif
Fig. 6 Energy dispersive X-ray (EDX) spectrum of Hymenodictyon orixense-synthesized silver nanoparticles showing the presence of different phyto-elements as capping agents.

As a general trend, the shape of plant-synthesized Ag NPs was spherical, with exception of some neem-synthesized Ag NPs. They are poly-disperse, with spherical or flat, plate-like, morphology, and mean size range of 5–35 nm in size.32 For example, Ag NPs fabricated using Emblica officinalis were also predominantly spherical with an average size of 16.8 nm ranging from 7.5 to 25 nm.33 Most of the Ag NPs was roughly circular in shape with smooth edges, as in the case of our H. orixense-fabricated Ag NPs. In agreement our these findings, Ag NPs from Annona squamosa leaf extract were spherical in shape with an average size ranging from 20 to 100 nm34 while Thirunavokkarasu et al.35 reported spherical nanoparticles with size ranging from 8 to 90 nm in Desmodium gangeticum. TEM images showed that the surface of the Ag NPs was often surrounded by a grey thin layer of some material, which might be due to the capping organic constituents of the plant broth, as highlighted by Rafiuddin.36

Larvicidal activity against malaria, chikungunya and Japanese encephalitis vectors

In laboratory conditions, the H. orixense aqueous leaf extract showed larvicidal activity against A. subpictus, A. albopictus and C. tritaeniorhynchus; LC50 values were 104.62, 113.88 and 123.55 μg ml−1, for A. subpictus, A. albopictus and C. tritaeniorhynchus, respectively (Table 1). Recently, a growing number of plant extracts have been found effective against C. quinquefasciatus larvae.37,38 Furthermore, the H. orixense-synthesized Ag NPs were highly toxic against A. subpictus, A. albopictus and C. tritaeniorhynchus larvae; LC50 values were 17.10, 18.74 and 20.08 μg ml−1, A. subpictus, A. albopictus and C. tritaeniorhynchus, respectively (Table 2). Controls treatments, including the bioassays testing Ag+ ions at the same concentrations of the tested Ag NP showed no mortality, in agreement with Jayaseelan et al.39 and Marimuthu et al.40 In latest years, a growing number of plant-synthesized Ag NP have been studied for their excellent larvicidal activity against important mosquito vectors.9,10 For instance, comparable toxicity rates have been recently reported for Ag NPs synthesized using Chomelia asiatica against A. stephensi larvae (LC50 = 17.95 ppm).12 The mortality effect evoked by Ag NPs on mosquito larvae and pupae may be due by the small size of the Ag NPs, which allows their passage through the insect cuticle and into individual cells, where they interfere with moulting and other physiological processes.9 The residual toxicity of silver ions against mosquito larvae covered a little role, since the peak in Fig. 1b was saturated after 180 min, indicating complete reduction of silver nitrate.41
Table 1 Larvicidal activity of the Hymenodictyon orixense aqueous leaf extract against the mosquito vectors Anopheles subpictus, Aedes albopictus and Culex tritaeniorhynchus
Mosquito species Concentration (μg ml−1) Mortality (%) ± SDa LC50 (μg ml−1) (LCL-UCL) LC90 (μg ml−1) (LCL-UCL) Slope Regression equation χ2 (d.f.)
a Values are mean ± SD of five replicates. No mortality was observed in the control. SD = standard deviation. LC50 = lethal concentration that kills 50% of the exposed organisms. LC90 = lethal concentration that kills 90% of the exposed organisms. UCL = 95% upper confidence limit. LCL = 95% lower confidence limit. χ2 = chi square. d.f. = degrees of freedom. n.s. = not significant (α = 0.05).
A. subpictus 50 27.5 ± 0.4 104.62 (92.82–115.08) 205.84 (190.92–225.46) 3.18 y = 10.45 + 0.368x 2.950 (4) n.s.
100 46.2 ± 0.8
150 68.4 ± 1.2
200 87.3 ± 0.6
250 99.0 ± 0.8
A. albopictus 50 24.2 ± 0.6 113.88 (102.25–124.42) 219.24 (203.41–240.11) 2.85 y = 6.11 + 0.375x 1.872 (4) n.s.
100 42.6 ± 1.2
150 63.7 ± 0.8
200 84.1 ± 0.4
250 97.2 ± 0.8
C. tritaeniorhynchus 50 21.4 ± 1.2 123.55 (111.97–134.27) 233.68 (216.64–256.31) 2.65 y = 2.49 + 0.376x 1.349 (4) n.s.
100 38.6 ± 0.6
150 60.2 ± 0.8
200 79.3 ± 0.4
250 95.1 ± 0.6


Table 2 Larvicidal activity of silver nanoparticles synthesized using the Hymenodictyon orixense leaf extract against the mosquito vectors Anopheles subpictus, Aedes albopictus and Culex tritaeniorhynchus
Mosquito species Concentration (μg ml−1) Mortality (%) ± SDa LC50 (μg ml−1) (LCL-UCL) LC90 (μg ml−1) (LCL-UCL) Slope Regression equation χ2 (d.f.)
a Values are mean ± SD of five replicates. No mortality was observed in the control. SD = standard deviation. LC50 = lethal concentration that kills 50% of the exposed organisms. LC90 = lethal concentration that kills 90% of the exposed organisms. UCL = 95% upper confidence limit. LCL = 95% lower confidence limit. χ2 = chi square. d.f. = degrees of freedom. n.s. = not significant (α = 0.05).
A. subpictus 8 25.8 ± 1.2 17.10 (15.26–18.75) 33.16 (30.78–36.27) 2.95 y = 8.94 + 2.335x 3.185 (4) n.s.
16 46.3 ± 0.8
24 67.2 ± 0.6
32 86.5 ± 0.4
40 99.1 ± 0.8
A. albopictus 8 22.4 ± 0.4 18.74 (16.95–20.39) 35.25 (32.75–38.54) 2.59 y = 3.89 + 2.394x 3.621 (4) n.s.
16 41.7 ± 1.2
24 62.8 ± 0.6
32 81.6 ± 0.8
40 98.2 ± 0.6
C. tritaeniorhynchus 8 19.3 ± 0.6 20.08 (18.32–21.74) 36.94 (34.33–40.38) 2.39 y = 0.2 + 2.425x 2.708 (4) n.s.
16 38.6 ± 0.4
24 59.4 ± 1.2
32 78.2 ± 0.6
40 96.5 ± 0.8


Biosafety on non-target mosquito predators

In our experiments, the toxicity treatments achieved negligible toxicity against D. indicus, with LC50 values ranging from 833.59 to 6233.15 μg ml−1 (Table 3). Focal observations highlighted that longevity the study species was not altered for 10 days after testing. SI indicated that H. orixense-fabricated Ag NP were less toxic to the non-target organism tested if compared to the targeted mosquitoes (Table 4). Currently, moderate knowledge is available about the acute toxicity of mosquitocidal nanoparticles towards non-target aquatic species.9 Recently, the genotoxic effect of silver nanoparticles synthesized using neem cake was studied on Carassius auratus using the comet assay and micronucleus frequency test. DNA damage was evaluated on peripheral erythrocytes sampled at different time intervals from the treatment. Interestingly, no significant damages were found at doses below 12 ppm.42 Furthermore, Spergularia rubra- and Pergularia daemia-synthesized Ag NP did not exhibit any evident toxicity effect against Poecilia reticulata fishes, after 48 h of exposure to LC50 and LC90 values calculated on IV instar larvae of A. aegypti and A. stephensi.43 Haldar et al.44 did not detected toxicity of Ag NP produced using dried green fruits of D. roxburghii against P. reticulata, after 48 h-exposure to LC50 of IV instar larvae of A. stephensi and C. quinquefasciatus. Rawani et al.45 showed that mosquitocidal Ag NP synthesized using Solanum nigrum berry extracts were not toxic against two mosquito predators, Toxorhynchites larvae and Diplonychus annulatum, and Chironomus circumdatus larvae, exposed to lethal concentrations of dry nanoparticles calculated on A. stephensi and C. quinquefasciatus larvae. Ag NPs biosynthesized using the 2,7-bis[2-[diethylamino]-ethoxy]fluorence isolate from the Melia azedarach leaves did not show acute toxicity against Mesocyclops pehpeiensis copepods.46 Interestingly, the exposure to extremely low doses (e.g. 1 ppm) of green-synthesized Ag NP did not negatively influence the predation efficiency of a number of mosquito predators of relevance for mosquito control.9,41
Table 3 Toxicity of Hymenodictyon orixense aqueous leaf extract and green-synthesized silver nanoparticles against the non-target mosquito predator Diplonychus indicus
Treatment Concentration (μg ml−1) Mortality (%) ± SDa LC50 (μg ml−1) (LCL-UCL) LC90 (μg ml−1) (LCL-UCL) Slope Regression equation χ2 (d.f.)
a Values are mean ± SD of five replicates. No mortality was observed in the control. SD = standard deviation. LC50 = lethal concentration that kills 50% of the exposed organisms. LC90 = lethal concentration that kills 90% of the exposed organisms. UCL = 95% upper confidence limit. LCL = 95% lower confidence limit. χ2 = chi square. d.f. = degrees of freedom. n.s. = not significant (α = 0.05).
Aqueous leaf extract 3000 26.2 ± 1.2 6233.15 (5563.16–6831.64) 11[thin space (1/6-em)]942.84 (11[thin space (1/6-em)]096.82–13[thin space (1/6-em)]047.17) 2.79 y = 9.67 + 0.006x 4.737 (4) n.s.
6000 48.3 ± 0.6
9000 67.5 ± 0.4
12[thin space (1/6-em)]000 89.4 ± 0.8
15[thin space (1/6-em)]000 100.0 ± 0.0
Silver nanoparticles 400 27.3 ± 0.4 833.59 (741.65–915.37) 1619.86 (1503.63–1772.15) 2.98 y = 10.15 + 0.047x 4.784 (4) n.s.
800 46.5 ± 0.8
1200 68.9 ± 1.2
1600 87.2 ± 0.6
2000 100.0 ± 0.0


Table 4 Suitability index of the non-target aquatic organism Diplonychus indicus over young instars of Anopheles subpictus, Aedes albopictus and Culex tritaeniorhynchus exposed to Hymenodictyon orixense aqueous leaf extract and green-synthesized silver nanoparticles
Treatment A. subpictus A. albopictus C. tritaeniorhynchus
Aqueous leaf extract 59.57 54.73 50.45
Silver nanoparticles 48.74 44.48 41.51


Conclusions

Overall, here we synthesized mosquitocidal silver nanoparticles using a cheap aqueous extract of H. orixense leaves as reducing and stabilizing agent. The bio-reduced Ag NPs were mostly spherical in shape, crystalline in nature, with face-cantered cubic geometry, and their mean size was 25–30 nm. This research highlighted that H. orixense-synthesized Ag NPs are easy to produce, stable over time, and can be employed at low dosages to strongly reduce populations of mosquito vectors, with little detrimental effects on the non-target mosquito predator D. indicus.

Conflicts of interest

The Authors declare no conflicts of interest.

Compliance with ethical standards

All applicable international and national guidelines for the care and use of animals were followed. All procedures performed in studies involving animals were in accordance with the ethical standards of the institution or practice at which the studies were conducted.

Acknowledgements

Dr Igor Medintz and two anonymous reviewers kindly improved an earlier version of the manuscript. The authors would like to thank Professor and Head, Department of Zoology, Annamalai University for the laboratory facilities provided. We also acknowledge the cooperation of staff members of the VCRC (ICMR), Pondicherry.

References

  1. G. Benelli and H. Mehlhorn, Parasitol. Res., 2016, 115(5), 1747–1754,  DOI:10.1007/s00436-016-4971-z.
  2. M. S. Klempner, T. R. Unnasch and L. T. Hu, N. Engl. J. Med., 2007, 356, 2567–2569 CrossRef CAS PubMed.
  3. H. Mehlhorn, K. A. Al-Rasheid, S. Al-Quraishy and F. Abdel-Ghaffar, Parasitol. Res., 2012, 110, 259–265 CrossRef PubMed.
  4. G. Benelli, Parasitol. Res., 2015, 114, 2801–2805 CrossRef PubMed.
  5. O. O. Anyaele and A. A. S. Amusan, Afr. J. Biomed. Res., 2003, 6, 49–53 Search PubMed.
  6. A. Amer and H. Mehlhorn, Parasitol. Res., 2006, 99, 466–472 CrossRef PubMed.
  7. G. Benelli, Parasitol. Res., 2015, 114, 3201–3212 CrossRef PubMed.
  8. R. Pavela, Ind. Crops Prod., 2015, 76, 174–187 CrossRef CAS.
  9. G. Benelli, Parasitol. Res., 2016, 115, 23–24 CrossRef PubMed.
  10. M. Govindarajan, in Nanoparticles in the Fight Against Parasites - Parasitology Research Monographs, ed. H. Mehlhorn, Springer International Publishing, Switzerland, 2016, pp. 99–153 Search PubMed.
  11. S. S. Shankar, A. Rai, A. Ahmad and M. Sastry, J. Colloid Interface Sci., 2004, 275, 496–502 CrossRef CAS PubMed.
  12. U. Muthukumaran, M. Govindarajan and M. Rajeswary, Parasitol. Res., 2015, 114, 989–999 CrossRef PubMed.
  13. K. Veerekumar, M. Govindarajan and M. Rajeswary, Parasitol. Res., 2013, 112, 4073–4085 CrossRef PubMed.
  14. U. Muthukumaran, M. Govindarajan and M. Rajeswary, Parasitol. Res., 2015, 114, 1817–1827 CrossRef PubMed.
  15. R. W. Raut, S. Kolekar Niranjan, R. Lakkakula Jaya, D. Mendhulkar Vijay and B. Kashid Sahebrao, Nano-Micro Lett., 2010, 2, 106–113 CrossRef CAS.
  16. G. Oberdorster, A. Maynard, K. Donaldson and V. Castranova, Part. Fibre Toxicol., 2005, 2, 8–43 CrossRef PubMed.
  17. A. Saxena, R. M. Tripathi and R. P. Singh, Dig. J. Nanomater. Biostruct., 2010, 5, 427–432 Search PubMed.
  18. A. Ghani, Medicinal Plants of Bangladesh with chemical constituents and uses, Bangladesh, 2nd edn, 2003 Search PubMed.
  19. R. P. Rastogi and B. N. Mehrotra, Compendium of Indian Medicinal Plants, Lucknow, 1st edn, 1993 Search PubMed.
  20. M. Govindarajan and G. Benelli, Parasitol. Res., 2016, 115, 925–935 CrossRef PubMed.
  21. World Health Organization, Guidelines for laboratory and field testing of mosquito larvicides, WHO/CDS/WHOPES/GCDPP/2005.13, Switzerland, 2005 Search PubMed.
  22. N. Sivagnaname and M. Kalyanasundaram, Mem. Inst. Oswaldo Cruz, 2004, 99, 115–118 CrossRef CAS PubMed.
  23. D. J. Finney, Probit analysis, Cambridge University Press, UK, 3rd edn, 1971 Search PubMed.
  24. P. G. Deo, S. B. Hasan and S. K. Majumdar, Int. Pest Control, 1988, 30, 118–129 CAS.
  25. B. P. Singh, B. J. Hatton, B. Singh, A. L. Cowie and A. Kathuria, J. Environ. Qual., 2010, 39, 1–12 CrossRef PubMed.
  26. M. Zargar, A. A. Hamid, F. A. Bakar, M. N. Shamsudin, K. Shameli, F. Jahanshiri and F. Farahani, Molecules, 2011, 6, 6667–6676 CrossRef PubMed.
  27. A. Ahmad, M. Mukherjee, D. Mandal, S. Senapati, M. I. Khan and R. Kumar, Colloids Surf., B, 2003, 28, 313–318 CrossRef CAS.
  28. N. N. Swe, Journal of the Myanmar Academy of Arts and Science, 2008, 6, 193–203 Search PubMed.
  29. N. Shanmugam, P. Rajkamal, S. Cholan, N. Kannadasan, K. Sathishkumar, G. Viruthagiri and A. Sundaramanickam, Appl. Nanosci., 2014, 4, 881–888 CrossRef CAS.
  30. S. Li, Y. Shen, A. Xie, X. Yu, L. Qiu, L. Zhang and Q. Zhang, Green Chem., 2007, 9, 852 RSC.
  31. P. Prakash, P. Gnanaprakasama, R. Emmanuel, S. Arokiyaraj and M. Saravanan, Colloids Surf., B, 2013, 108, 255–259 CrossRef CAS PubMed.
  32. S. S. Shankar, A. Rai, A. Ahmad and M. Sastry, Nanotechnol. Appl., 2004, 1, 69–77 Search PubMed.
  33. B. Ankamwar, M. Chaudhary and M. Sastry, Synth. React. Inorg., Met.-Org., Nano-Met. Chem., 2005, 35, 19–26 CrossRef CAS.
  34. R. Vivek, R. Thangam, K. Muthuchelian, P. Gunasekaran, K. Kaveri and S. Kannan, Process Biochem., 2012, 47, 2405–2410 CrossRef CAS.
  35. M. Thirunavokkarasu, U. Balaji, S. Behera, P. K. Panda and B. K. Mishra, Spectrochim. Acta, Part A, 2013, 116, 424–427 CrossRef PubMed.
  36. Z. Z. Rafiuddin, Colloids Surf., B, 2013, 108, 90–94 CrossRef PubMed.
  37. D. Kumar, R. Chawla, P. Dhamodaram and N. Balakrishnan, Parasitol. Res., 2014, 236838 Search PubMed.
  38. G. Chitra, G. Balasubramani, R. Ramkumar, R. Sowmiya and P. Perumal, Parasitol. Res., 2014, 114, 1407–1415 CrossRef PubMed.
  39. C. Jayaseelan, A. A. Rahuman, G. Rajakumar, T. Santhoshkumar, A. V. Kirthi, S. Marimuthu, A. Bagavan, C. Kamaraj, A. A. Zahir, G. Elango, K. Velayutham, K. V. B. Rao, L. Karthik and S. Raveendran, Parasitol. Res., 2012, 111, 921–933 CrossRef PubMed.
  40. S. Marimuthu, A. A. Rahuman, G. Rajakumar, T. Santhoshkumar, A. V. Kirthi, C. Jayaseelan, A. Bagavan, A. A. Zahir, G. Elango and C. Kamaraj, Parasitol. Res., 2011, 108, 1541–1549 CrossRef PubMed.
  41. K. Murugan, D. Dinesh, P. Jenil Kumar, C. Panneerselvam, J. Subramaniam, P. Madhiyazhagan, U. Suresh, M. Nicoletti, A. A. Alarfaj and M. A. Munusamy, Parasitol. Res., 2015, 114, 4645–4654 CrossRef PubMed.
  42. B. Chandramohan, K. Murugan, C. Panneerselvam, P. Madhiyazhagan, R. Chandirasekar, D. Dinesh, P. Mahesh Kumar, K. Kovendan, U. Suresh, J. Subramaniam, R. Rajaganesh, A. T. Aziz, B. Syuhei, M. Saleh Alsalhi, S. Devanesan, M. Nicoletti, H. Wei and G. Benelli, Parasitol. Res., 2016, 115, 1015–1025 CrossRef PubMed.
  43. C. D. Patil, H. P. Borase, S. V. Patil, R. B. Salunkhe and B. K. Salunke, Parasitol. Res., 2012, 111, 555–562 CrossRef PubMed.
  44. B. Haldar and G. Chandra, Parasitol. Res., 2013, 112, 1451–1459 CrossRef PubMed.
  45. A. Rawani, A. Ghosh and G. Chandra, Acta Trop., 2013, 128, 613–622 CrossRef CAS PubMed.
  46. R. Ramanibai and K. Velayutham, Res. Vet. Sci., 2015, 98, 82–88 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2016