[go: up one dir, main page]

 
 

Engineered Nanomaterials in the Environment

A special issue of Nanomaterials (ISSN 2079-4991).

Deadline for manuscript submissions: closed (31 January 2016) | Viewed by 54437

Special Issue Editors


E-Mail Website
Guest Editor
Commonwealth Scientific and Industrial Research Organization, Melbourne, Australia
Interests: environmental chemistry and toxicology; molecular environmental science; physicochemical factors and mechanisms controlling the transport and fate of inorganic and organic contaminants and manufactured nanomaterials within surface and subsurface environments; molecular-level controls on the bioavailability and toxicity of metals/nanomaterials; ecotoxicology and eco-genomics; biogeochemistry; critical zone science; climate science; nuclear waste management and disposal

E-Mail
Guest Editor
Division of Land and Water, CSIRO, Waite Campus, Urrbrae, SA 5064, Australia
Interests: soil chemistry; soil microbiology; critical zone science; soil colloids; ecotoxicology and emerging contaminants

Special Issue Information

Dear Colleagues,

Engineered nanomaterials (ENMs) are being incorporated into a rapidly increasing number of consumer products including sunscreens, cosmetics, pharmaceuticals, and textiles. The ENMs in consumer products can be released into waste streams during use and will concentrate in sewage sludge during wastewater treatment. As a result, estimated concentrations of ENMs within sludge and the biosolids produced from them have increased dramatically over the past few years. In regions where biosolids are used as fertilizer, land application of biosolids is a significant pathway by which ENMs will be introduced into agroecosystems. Furthermore, the potential to develop ENM-based pesticides and fertilizers is currently being explored. These agricultural applications present the possibility of widespread, large volume discharge of ENMs into agroecosystems. As questions remain over the potential environmental impact of ENMs, scientific and community concerns persist over their human and environmental safety. In this Special Issue of Nanomaterials, we present recent research findings and ideas concerning our understanding of the fate, bioavailability, and potential toxicity of ENMs in the environment.

Prof. Dr. Paul Bertsch
Dr. Jonathan Judy
Guest Editors

Manuscript Submission Information

Manuscripts should be submitted online at www.mdpi.com by registering and logging in to this website. Once you are registered, click here to go to the submission form. Manuscripts can be submitted until the deadline. All submissions that pass pre-check are peer-reviewed. Accepted papers will be published continuously in the journal (as soon as accepted) and will be listed together on the special issue website. Research articles, review articles as well as short communications are invited. For planned papers, a title and short abstract (about 100 words) can be sent to the Editorial Office for announcement on this website.

Submitted manuscripts should not have been published previously, nor be under consideration for publication elsewhere (except conference proceedings papers). All manuscripts are thoroughly refereed through a single-blind peer-review process. A guide for authors and other relevant information for submission of manuscripts is available on the Instructions for Authors page. Nanomaterials is an international peer-reviewed open access semimonthly journal published by MDPI.

Please visit the Instructions for Authors page before submitting a manuscript. The Article Processing Charge (APC) for publication in this open access journal is 2900 CHF (Swiss Francs). Submitted papers should be well formatted and use good English. Authors may use MDPI's English editing service prior to publication or during author revisions.


Keywords

  • nanotoxicology
  • nanoparticles
  • nanomaterials
  • nanotechnology
  • nanopesticide
  • nanofertilizer
  • ecotoxicology

Benefits of Publishing in a Special Issue

  • Ease of navigation: Grouping papers by topic helps scholars navigate broad scope journals more efficiently.
  • Greater discoverability: Special Issues support the reach and impact of scientific research. Articles in Special Issues are more discoverable and cited more frequently.
  • Expansion of research network: Special Issues facilitate connections among authors, fostering scientific collaborations.
  • External promotion: Articles in Special Issues are often promoted through the journal's social media, increasing their visibility.
  • e-Book format: Special Issues with more than 10 articles can be published as dedicated e-books, ensuring wide and rapid dissemination.

Further information on MDPI's Special Issue polices can be found here.

Published Papers (8 papers)

Order results
Result details
Select all
Export citation of selected articles as:

Editorial

Jump to: Research

150 KiB  
Editorial
Engineered Nanomaterials in the Environment
by Jonathan D. Judy and Paul Bertsch
Nanomaterials 2016, 6(6), 106; https://doi.org/10.3390/nano6060106 - 6 Jun 2016
Cited by 3 | Viewed by 3870
Abstract
This Special Issue of Nanomaterials, “Engineered Nanomaterials in the Environment”, is comprised of one communication and five research articles.[...] Full article
(This article belongs to the Special Issue Engineered Nanomaterials in the Environment)

Research

Jump to: Editorial

3859 KiB  
Article
Separation of Bacteria, Protozoa and Carbon Nanotubes by Density Gradient Centrifugation
by Monika Mortimer, Elijah J. Petersen, Bruce A. Buchholz and Patricia A. Holden
Nanomaterials 2016, 6(10), 181; https://doi.org/10.3390/nano6100181 - 12 Oct 2016
Cited by 22 | Viewed by 10034
Abstract
Sustainable production and use of carbon nanotube (CNT)-enabled materials require efficient assessment of CNT environmental hazards, including the potential for CNT bioaccumulation and biomagnification in environmental receptors. Microbes, as abundant organisms responsible for nutrient cycling in soil and water, are important ecological receptors [...] Read more.
Sustainable production and use of carbon nanotube (CNT)-enabled materials require efficient assessment of CNT environmental hazards, including the potential for CNT bioaccumulation and biomagnification in environmental receptors. Microbes, as abundant organisms responsible for nutrient cycling in soil and water, are important ecological receptors for studying the effects of CNTs. Quantification of CNT association with microbial cells requires efficient separation of CNT-associated cells from individually dispersed CNTs and CNT agglomerates. Here, we designed, optimized, and demonstrated procedures for separating bacteria (Pseudomonas aeruginosa) from unbound multiwall carbon nanotubes (MWCNTs) and MWCNT agglomerates using sucrose density gradient centrifugation. We demonstrate separation of protozoa (Tetrahymena thermophila) from MWCNTs, bacterial agglomerates, and protozoan fecal pellets by centrifugation in an iodixanol solution. The presence of MWCNTs in the density gradients after centrifugation was determined by quantification of 14C-labeled MWCNTs; the recovery of microbes from the density gradient media was confirmed by optical microscopy. Protozoan intracellular contents of MWCNTs and of bacteria were also unaffected by the designed separation process. The optimized methods contribute to improved efficiency and accuracy in quantifying MWCNT association with bacteria and MWCNT accumulation in protozoan cells, thus supporting improved assessment of CNT bioaccumulation. Full article
(This article belongs to the Special Issue Engineered Nanomaterials in the Environment)
Show Figures

Figure 1

Figure 1
<p>Representative phase contrast images of <span class="html-italic">P. aeruginosa</span> incubated with 5 mg/L MWCNTs for 1 h: (<b>a</b>) pelleted bacteria after differential centrifugation, white arrow points to MWCNT agglomerate; and (<b>b</b>) bacteria after density gradient centrifugation in sucrose.</p>
Full article ">Figure 2
<p>Natural log-transformed <span class="html-italic">T. thermophila</span> cell counts per milliliter over 22 h growth with control or MWCNT-treated <span class="html-italic">P. aeruginosa</span> as a food source. Data points are the average of three replicates, and error bars indicate standard deviation values. Letters designate growth phases: L—lag phase, E—exponential phase, LE—late exponential phase, and S—stationary phase.</p>
Full article ">Figure 3
<p>Nomarski images of <span class="html-italic">T. thermophila</span> grown with MWCNT-encrusted bacteria (<b>a</b>,<b>b</b>) and control bacteria (<b>c</b>,<b>d</b>), taken before (<b>a</b>,<b>c</b>) and after (<b>b</b>,<b>d</b>) separation steps (differential centrifugation and density gradient centrifugations in iodixanol) at different growth phases. The round shapes inside protozoan cells are food vacuoles filled with bacteria. White arrows indicate bacteria provided as food source and black arrows show protozoan fecal pellets, i.e., secreted food vacuoles containing digested or partly digested bacteria.</p>
Full article ">Figure 4
<p>Representative phase contrast images of the pelleted fraction after centrifugation of <span class="html-italic">T. thermophila</span> suspensions through 10% iodixanol. The two images serve as replicates and are both shown since bacterial agglomerates (<b>left</b> image) and free bacterial cells (<b>right</b> image) were each observed. White arrows indicate bacterial agglomerates or cells that pelleted with the protozoan cells.</p>
Full article ">Figure 5
<p>Protozoan cell recovery percentages after density gradient centrifugation steps in iodixanol when sampled at four different protozoan growth phases. Protozoa were grown with control (untreated, white bars) or MWCNT-treated bacteria (blue bars). Data bars represent the mean values of three measurements and error bars are standard deviations. The values of “Number of protozoan cells before centrifugation” denote the total protozoan cell number in the concentrated samples after differential centrifugation but before density gradient centrifugation in iodixanol. Asterisks indicate significant difference from the recovery percentage at lag phase, both in control and MWCNT-amended exposure, using Tukey’s multiple comparisons test across all eight conditions, <span class="html-italic">p</span> ≤ 0.05 (**) and <span class="html-italic">p</span> ≤ 0.1 (*). There was no significant difference between the percentages of recovered protozoan cells grown with control bacteria (white bars) versus protozoan cells grown with MWCNT-encrusted bacteria (blue bars) within each growth phase.</p>
Full article ">Figure 6
<p>(<b>a</b>) Comparison of food vacuole numbers in protozoan cells before and after differential and density gradient centrifugations; and (<b>b</b>) correlation between the food vacuole numbers and MWCNT mass per protozoan cell when sampled at four different growth phases (indicated by the letters: L—lag phase, E—exponential phase, LE—late exponential phase and S—stationary phase). Numbers at each data point in panel (<b>a</b>) indicate mean food vacuole numbers before and after centrifugation (<span class="html-italic">x</span>; <span class="html-italic">y</span>), which were not statistically different based on the two-sample <span class="html-italic">t</span>-test (<span class="html-italic">p</span> ≤ 0.05) for each of the growth phases. Food vacuole numbers per cell are the mean values of 5 to 11 individual cells. MWCNT masses are based on the mean values of three LSC measurements of <sup>14</sup>C associated with radiolabeled MWCNTs. Error bars indicate standard deviations.</p>
Full article ">Scheme 1
<p>Schematic illustration of observed separation results for bacteria and protozoa after either bacterial exposure to MWCNTs, or protozoan exposure to bacterial prey with their cell-associated MWCNTs: (<b>a</b>) separation of MWCNT-associated bacteria (shown in red) from MWCNTs; and (<b>b</b>) separation of MWCNTs, bacteria and fecal pellets from protozoa (shown in yellow) that were grown with MWCNT-encrusted bacteria; free MWCNTs are included as a component of the system to account for possible “shedding” of MWCNTs from bacteria before or after the uptake into protozoan food vacuoles. Blue lines denote liquid levels in the tubes and interfaces of different liquids. FA—fixed angle rotor, SW—swinging bucket rotor (<a href="#app1-nanomaterials-06-00181" class="html-app">Table S2</a>).</p>
Full article ">
3459 KiB  
Article
Aggregation and Colloidal Stability of Commercially Available Al2O3 Nanoparticles in Aqueous Environments
by Julie Mui, Jennifer Ngo and Bojeong Kim
Nanomaterials 2016, 6(5), 90; https://doi.org/10.3390/nano6050090 - 13 May 2016
Cited by 52 | Viewed by 8202
Abstract
The aggregation and colloidal stability of three, commercially-available, gamma-aluminum oxide nanoparticles (γ-Al2O3 NPs) (nominally 5, 10, and 20–30 nm) were systematically examined as a function of pH, ionic strength, humic acid (HA) or clay minerals (e.g., montmorillonite) concentration using dynamic [...] Read more.
The aggregation and colloidal stability of three, commercially-available, gamma-aluminum oxide nanoparticles (γ-Al2O3 NPs) (nominally 5, 10, and 20–30 nm) were systematically examined as a function of pH, ionic strength, humic acid (HA) or clay minerals (e.g., montmorillonite) concentration using dynamic light scattering and transmission electron microscopy techniques. NPs possess pH-dependent surface charges, with a point of zero charge (PZC) of pH 7.5 to 8. When pH < PZC, γ-Al2O3 NPs are colloidally stable up to 100 mM NaCl and 30 mM CaCl2. However, significant aggregation of NPs is pronounced in both electrolytes at high ionic strength. In mixed systems, both HA and montmorillonite enhance NP colloidal stability through electrostatic interactions and steric hindrance when pH ≤ PZC, whereas their surface interactions are quite limited when pH > PZC. Even when pH approximates PZC, NPs became stable at a HA concentration of 1 mg·L−1. The magnitude of interactions and dominant sites of interaction (basal planes versus edge sites) are significantly dependent on pH because both NPs and montmorillonite have pH-dependent (conditional) surface charges. Thus, solution pH, ionic strength, and the presence of natural colloids greatly modify the surface conditions of commercial γ-Al2O3 NPs, affecting aggregation and colloidal stability significantly in the aqueous environment. Full article
(This article belongs to the Special Issue Engineered Nanomaterials in the Environment)
Show Figures

Figure 1

Figure 1
<p>X-ray diffraction (XRD) patterns of commercial γ-Al<sub>2</sub>O<sub>3</sub> nanoparticles (NPs) in size of: (<b>a</b>) 5 nm; (<b>b</b>) 10 nm; and (<b>c</b>) 20–30 nm. Note that 20–30 nm γ-Al<sub>2</sub>O<sub>3</sub> NPs have a trace amount of boehmite (γ-AlOOH) (PDF 04-010-5683) marked with *, as an impurity.</p>
Full article ">Figure 2
<p>Bright field transmission electron microscopy (TEM) images of γ-Al<sub>2</sub>O<sub>3</sub> NPs in size of: (<b>a</b>) 5 nm; (<b>b</b>) 10 nm; and (<b>c</b>) 20–30 nm.</p>
Full article ">Figure 3
<p>Z-average (Z<sub>ave</sub>) hydrodynamic diameter (○) and ζ potential (●) of: (<b>a</b>) 5 nm; (<b>b</b>) 10 nm; and (<b>c</b>) 20–30 nm γ-Al<sub>2</sub>O<sub>3</sub> NP aggregates, as a function of pH.</p>
Full article ">Figure 4
<p>(<b>a</b>) Z<sub>ave</sub> hydrodynamic diameter of 5 nm (□), 10 nm (○), and 20–30 nm (△) γ-Al<sub>2</sub>O<sub>3</sub> NP aggregates, as a function of ionic strength (NaCl); (<b>b</b>) Z<sub>ave</sub> hydrodynamic diameter of 5 nm (□), 10 nm (○), and 20–30 nm (△) γ-Al<sub>2</sub>O<sub>3</sub> NP aggregates, as a function of ionic strength (CaCl<sub>2</sub>); (<b>c</b>) ζ potential of 5 nm (■), 10 nm (●), and 20–30 nm (▲) γ-Al<sub>2</sub>O<sub>3</sub> NP aggregates, as a function of ionic strength (NaCl); and (<b>d</b>) ζ potential of 5 nm (■), 10 nm (●), and 20–30 nm (▲) γ-Al<sub>2</sub>O<sub>3</sub> NP aggregates, as a function of ionic strength (CaCl<sub>2</sub>).</p>
Full article ">Figure 5
<p>(<b>a</b>) Z<sub>ave</sub> hydrodynamic diameter (○) and ζ potential (●) of 10 nm of γ-Al<sub>2</sub>O<sub>3</sub> NPs as a function of HA concentration at pH close to the point of zero charge (PZC) of Al<sub>2</sub>O<sub>3</sub>; and (<b>b</b>) bright field TEM images of 10 nm γ-Al<sub>2</sub>O<sub>3</sub> NPs with 10 mg·L<sup>−1</sup> of HA at pH close to PZC of Al<sub>2</sub>O<sub>3</sub>.</p>
Full article ">Figure 6
<p>Bright field TEM images of 10 nm γ-Al<sub>2</sub>O<sub>3</sub> NPs with montmorillonite at pH &lt; PZC of Al<sub>2</sub>O<sub>3</sub> NPs.</p>
Full article ">Figure 7
<p>Bright field TEM images of 10 nm γ-Al<sub>2</sub>O<sub>3</sub> NPs with montmorillonite at PZC of Al<sub>2</sub>O<sub>3</sub> NPs.</p>
Full article ">Figure 8
<p>Bright field TEM images of 10 nm γ-Al<sub>2</sub>O<sub>3</sub> NPs with montmorillonite at pH &gt; PZC of Al<sub>2</sub>O<sub>3</sub> NPs.</p>
Full article ">
1390 KiB  
Communication
Gold Nanomaterial Uptake from Soil Is Not Increased by Arbuscular Mycorrhizal Colonization of Solanum Lycopersicum (Tomato)
by Jonathan D. Judy, Jason K. Kirby, Mike J. McLaughlin, Timothy Cavagnaro and Paul M. Bertsch
Nanomaterials 2016, 6(4), 68; https://doi.org/10.3390/nano6040068 - 13 Apr 2016
Cited by 10 | Viewed by 4841
Abstract
Bioaccumulation of engineered nanomaterials (ENMs) by plants has been demonstrated in numerous studies over the past 5–10 years. However, the overwhelming majority of these studies were conducted using hydroponic systems and the degree to which the addition of the biological and chemical components [...] Read more.
Bioaccumulation of engineered nanomaterials (ENMs) by plants has been demonstrated in numerous studies over the past 5–10 years. However, the overwhelming majority of these studies were conducted using hydroponic systems and the degree to which the addition of the biological and chemical components present in the soil might fundamentally alter the potential of plant bioaccumulation of ENMs is unclear. Here, we used two genotypes of Solanum lycopersicum (tomato), reduced mycorrhizal colonization (rmc), a mutant which does not allow arbuscular mycorrhizal fungi (AMF) colonization, and its progenitor, 76R, to examine how colonization by AMF alters trends of gold ENM bioaccumulation from a natural soil. Gold was taken up and bioaccumulated by plants of both genotypes. Gold concentrations were significantly higher in the rmc treatment although this was likely attributable to the large differences in biomass between the 76R and rmc plants. Regardless, there was little evidence that AMF played a significant role in trafficking Au ENMs into the plants. Furthermore, despite very low NH4NO3 extractable Au concentrations, Au accumulated at the root-soil interface. Although this observation would seem to suggest that ENMs may have potential to influence this particularly biologically active and important soil compartment, we observed no evidence of this here, as the 76R plants developed a robust AMF symbiosis despite accumulation of Au ENMs at the rhizoplane. Full article
(This article belongs to the Special Issue Engineered Nanomaterials in the Environment)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Dry shoot biomass (<b>A</b>); mycorrhizal colonization frequency (<b>B</b>); shoot Au concentrations (<b>C</b>); and shoot Au uptake (<b>D</b>) measured in 76R and <span class="html-italic">rmc</span> tomato plants. Error bars represent standard deviation. The * indicates a significant difference at α ≤ 0.01 as determined by a either a 2-sided <span class="html-italic">T</span>-test (concentration and uptake) or a Mann-Whitney <span class="html-italic">U</span>-test (biomass and colonization).</p>
Full article ">Figure 2
<p>Micrographs (<b>left</b>) and laser ablation inductively-coupled plasma mass spectrometry (LA-ICP-MS) maps (<b>right</b>) of root cross-sections collected from root samples from (<b>top</b>) 76R and (<b>bottom</b>) <span class="html-italic">rmc</span> tomato plants. Color bars inset in LA-ICP-MS maps show relationship between counts per second (CPS) and color for each map. Ep = epidermis. En = endodermis. Lr = lateral root.</p>
Full article ">Figure 3
<p>TEM (transmission electron microscopy) micrograph and energy-dispersive X-ray spectroscopy (EDS) spectrum characterizing gold engineered nanomaterials (ENMs). Copper detected is result of the use of Cu TEM grids. cps: counts per second.</p>
Full article ">
2450 KiB  
Article
Nanoparticles Composed of Zn and ZnO Inhibit Peronospora tabacina Spore Germination in vitro and P. tabacina Infectivity on Tobacco Leaves
by George Wagner, Victor Korenkov, Jonathan D. Judy and Paul M. Bertsch
Nanomaterials 2016, 6(3), 50; https://doi.org/10.3390/nano6030050 - 16 Mar 2016
Cited by 58 | Viewed by 6069
Abstract
Manufactured nanoparticles (NPs) are increasingly being used for commercial purposes and certain NP types have been shown to have broad spectrum antibacterial activity. In contrast, their activities against fungi and fungi-like oomycetes are less studied. Here, we examined the potential of two types [...] Read more.
Manufactured nanoparticles (NPs) are increasingly being used for commercial purposes and certain NP types have been shown to have broad spectrum antibacterial activity. In contrast, their activities against fungi and fungi-like oomycetes are less studied. Here, we examined the potential of two types of commercially available Zn NPs (Zn NPs and ZnO NPs) to inhibit spore germination and infectivity on tobacco leaves resulting from exposure to the fungi-like oomycete pathogen Peronospora tabacina (P. tabacina). Both types of NPs, as well as ZnCl2 and bulk ZnO control treatments, inhibited spore germination compared to a blank control. ZnO ENMs were shown to be a much more powerful suppressor of spore germination and infectivity than bulk ZnO. ZnO and Zn NPs significantly inhibited leaf infection at 8 and 10 mg·L−1, respectively. Both types of NPs were found to provide substantially higher concentration dependent inhibition of spore germination and infectivity than could be readily explained by the presence of dissolved Zn. These results suggest that both NP types have potential for use as economic, low-dose, potentially non-persistent anti-microbial agents against the oomycete P. tabacina. Full article
(This article belongs to the Special Issue Engineered Nanomaterials in the Environment)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Transmission electron microscope (TEM) micrographs of (<b>A</b>) Zn nanoparticles (NPs) and (<b>B</b>) ZnO NPs.</p>
Full article ">Figure 2
<p>Dissolved Zn measured in the background of treatment suspensions (<b>top</b>) and the percentage of total Zn that is present at dissolved Zn (<b>bottom</b>).</p>
Full article ">Figure 3
<p>Germination frequency of <span class="html-italic">Peronospora tabacina</span> (<span class="html-italic">P. tabacina</span>) spores in the presence of Zn treatments. Error bars = standard deviation.</p>
Full article ">Figure 4
<p>Effects on germination tube length in the presence of ZnO nanoparticles (NPs). Micrographs show spores germinated in water (<b>A</b>), in 16 mg·L<sup>−1</sup> ZnO NPs (<b>B</b>), and 52 mg·L<sup>−1</sup> ZnO NPs (<b>C</b>). Note the haloed appearance and long germination tubes of spores germinated in water <span class="html-italic">versus</span> the dark nature of most of those germinated in 16 mg·L<sup>−1</sup> ZnO NPs, and all those germinated in 52 mg·L<sup>−1</sup> ZnO NPs. Also note in <a href="#nanomaterials-06-00050-f004" class="html-fig">Figure 4</a>B that some spores had a haloed appearance but these only produced short germination tubes.</p>
Full article ">Figure 5
<p>Spore germination tube length as measured for selected treatments. Error bars = standard deviation.</p>
Full article ">Figure 6
<p>Percentage of tobacco plants infected by <span class="html-italic">P. tabacina</span> in the presence of Zn treatments. Error bars = standard deviation.</p>
Full article ">
2095 KiB  
Article
Toxicity Testing of Pristine and Aged Silver Nanoparticles in Real Wastewaters Using Bioluminescent Pseudomonas putida
by Florian Mallevre, Camille Alba, Craig Milne, Simon Gillespie, Teresa F. Fernandes and Thomas J. Aspray
Nanomaterials 2016, 6(3), 49; https://doi.org/10.3390/nano6030049 - 11 Mar 2016
Cited by 30 | Viewed by 5385
Abstract
Impact of aging on nanoparticle toxicity in real matrices is scarcely investigated due to a lack of suitable methodologies. Herein, the toxicity of pristine and aged silver nanoparticles (Ag NPs) to a bioluminescent Pseudomonas putida bioreporter was measured in spiked crude and final [...] Read more.
Impact of aging on nanoparticle toxicity in real matrices is scarcely investigated due to a lack of suitable methodologies. Herein, the toxicity of pristine and aged silver nanoparticles (Ag NPs) to a bioluminescent Pseudomonas putida bioreporter was measured in spiked crude and final wastewater samples (CWs and FWs, respectively) collected from four wastewater treatment plants (WWTPs). Results showed lower toxicity of pristine Ag NPs in CWs than in FWs. The effect of the matrix on the eventual Ag NP toxicity was related to multiple physico-chemical parameters (biological oxygen demand (BOD), chemical oxygen demand (COD), total suspended solids (TSS) pH, ammonia, sulfide and chloride) based on a multivariate analysis. However, no collection site effect was concluded. Aged Ag NPs (up to eight weeks) were found less toxic than pristine Ag NPs in CWs; evident increased aggregation and decreased dissolution were associated with aging. However, Ag NPs exhibited consistent toxicity in FWs despite aging; comparable results were obtained in artificial wastewater (AW) simulating effluent. The study demonstrates the potency of performing nanoparticle acute toxicity testing in real and complex matrices such as wastewaters using relevant bacterial bioreporters. Full article
(This article belongs to the Special Issue Engineered Nanomaterials in the Environment)
Show Figures

Figure 1

Figure 1
<p>Real time monitoring of silver nanoparticle (Ag NP) toxicity in wastewaters. Relative luminescence output evolutions over time by <span class="html-italic">Pseudomonas putida</span> (<span class="html-italic">P. putida</span>) BS566::luxCDABE when challenged up to 200 mg·L<sup>−1</sup> of Ag NPs in (<b>a</b>) crude and (<b>b</b>) final wastewaters from Site 2 are shown. Four out of the nine used NP concentrations are plotted; the entire graphics as well as results with samples from other sites are presented in <a href="#app1-nanomaterials-06-00049" class="html-app">Supplementary Materials (Figures S1–S4)</a>. Background signal from used matrices and effect of Ag dispersant (at 50 mg·L<sup>−1</sup>) are also presented. Data are mean ± standard error of the mean (SEM) (<span class="html-italic">n</span> = 4).</p>
Full article ">Figure 2
<p>Derived toxicity values. Toxicity results were plotted as (response) = <span class="html-italic">f</span>(log[Ag NPs]) for selected time points and IC<sub>50</sub> values (half maximal inhibitory concentrations) were derived by fitting a four parameter concentration–response model. An example of obtained fits for one test with crude and final wastewaters (CWs and FWs, respectively) from Site 2 (including four replicates <span class="html-italic">per</span> condition) is shown in (<b>a</b>). The comparison of all calculated IC<sub>50</sub> values at 1 h is shown in (<b>b</b>). Data are mean ± SEM (<span class="html-italic">n</span> = 4), significant differences are represented by with <span class="html-italic">p</span> &lt; 0.1 (*) or <span class="html-italic">p</span> &lt; 0.05 (**) following analysis with unpaired t-tests. Derived IC<sub>50</sub> values at 0.5 and 2 h are presented in <a href="#app1-nanomaterials-06-00049" class="html-app">Supplementary Materials (Figure S5)</a>.</p>
Full article ">Figure 3
<p>Results from multivariate analysis. Ordination diagram of thirty two wastewater samples from four wastewater treatment plants (WWTPs) (Site 1 in blue, Site 2 in red, Site 3 in orange and Site 4 in green) obtained by canonical correlation analysis considering height biochemical parameters (BOD, COD, TSS, ammonia, pH, chloride, sulfide and total plate count) as environmental variables constrained by one explanatory variable, the derived IC<sub>50</sub> values at 1 h.</p>
Full article ">Figure 4
<p>Ag NP characterization in wastewaters. Ag NPs at 10 mg·L<sup>−1</sup> in crude and final wastewaters (CWs and FWs, respectively) were characterized by dynamic light scattering (DLS) and ultraviolet-visible spectroscopy (UV-vis). The hydrodynamic size results are shown in (<b>a</b>). The zeta potential data are plotted in (<b>b</b>). Data are in both cases mean ± SEM (<span class="html-italic">n</span> = 4), significantly different by unpaired t-test with <span class="html-italic">p</span> &lt; 0.05 (**). In (<b>c</b>) is shown an example of typical spectra of absorbance (between 300 and 900 nm) obtained for spiked FWs and CWs from Site 2.</p>
Full article ">Figure 5
<p>Effects of aging. Fate and toxicity of Ag NPs were tested after 0, 1, 2, 4 and 8 weeks of aging in FW and CW from Site 3. Derived IC<sub>50</sub> values at 1 h from ecotoxicity assays are presented in (<b>a</b>). Absorbance spectra obtained by UV-vis with Ag NPs at 10 mg·L<sup>−1</sup> are presented in (<b>b</b>). Hydrodynamic size and zeta potential information, determined by DLS (Ag NPs at 10 mg·L<sup>−1</sup>), are presented in (<b>c</b>) and (<b>d</b>) respectively. Data are mean ± SEM (<span class="html-italic">n</span> = 3), significant differences are represented as <span class="html-italic">p</span> &lt; 0.1 (*) or <span class="html-italic">p</span> &lt; 0.05 (**) following analysis with unpaired t-tests. Corresponding results obtained in artificial wastewater (AW) are presented in <a href="#app1-nanomaterials-06-00049" class="html-app">Supplementary Materials (Figure S7)</a>.</p>
Full article ">
551 KiB  
Article
EU Regulation of Nanobiocides: Challenges in Implementing the Biocidal Product Regulation (BPR)
by Anna Brinch, Steffen Foss Hansen, Nanna B. Hartmann and Anders Baun
Nanomaterials 2016, 6(2), 33; https://doi.org/10.3390/nano6020033 - 16 Feb 2016
Cited by 41 | Viewed by 9121
Abstract
The Biocidal Products Regulation (BPR) contains several provisions for nanomaterials (NMs) and is the first regulation in the European Union to require specific testing and risk assessment for the NM form of a biocidal substance as a part of the information requirements. Ecotoxicological [...] Read more.
The Biocidal Products Regulation (BPR) contains several provisions for nanomaterials (NMs) and is the first regulation in the European Union to require specific testing and risk assessment for the NM form of a biocidal substance as a part of the information requirements. Ecotoxicological data are one of the pillars of the information requirements in the BPR, but there are currently no standard test guidelines for the ecotoxicity testing of NMs. The overall objective of this work was to investigate the implications of the introduction of nano-specific testing requirements in the BPR and to explore how these might be fulfilled in the case of copper oxide nanoparticles. While there is information and data available in the open literature that could be used to fulfill the BPR information requirements, most of the studies do not take the Organisation for Economic Co-operation and Development’s nanospecific test guidelines into consideration. This makes it difficult for companies as well as regulators to fulfill the BPR information requirements for nanomaterials. In order to enable a nanospecific risk assessment, best practices need to be developed regarding stock suspension preparation and characterization, exposure suspensions preparation, and for conducting ecotoxicological test. Full article
(This article belongs to the Special Issue Engineered Nanomaterials in the Environment)
Show Figures

Figure 1

Figure 1
<p>Number of studies potentially fulfilling the Biocidal Product Regulation (BPR) information requirements for ecotoxicity tests.</p>
Full article ">Figure 2
<p>Number of ecotoxicity studies on copper oxide nanoparticles considering the reporting and characterization parameters recommended in the Organisation for Economic Co-operation and Development (OECD) guidance document [<a href="#B18-nanomaterials-06-00033" class="html-bibr">18</a>].</p>
Full article ">
1087 KiB  
Article
Organic Phase Change Nanoparticles for in-Product Labeling of Agrochemicals
by Miao Wang, Binh Duong and Ming Su
Nanomaterials 2015, 5(4), 1810-1819; https://doi.org/10.3390/nano5041810 - 28 Oct 2015
Cited by 11 | Viewed by 6102
Abstract
There is an urgent need to develop in-product covert barcodes for anti-counterfeiting of agrochemicals. This paper reports a new organic nanoparticle-based in-product barcode system, in which a panel of organic phase change nanoparticles is added as a barcode into in a variety of [...] Read more.
There is an urgent need to develop in-product covert barcodes for anti-counterfeiting of agrochemicals. This paper reports a new organic nanoparticle-based in-product barcode system, in which a panel of organic phase change nanoparticles is added as a barcode into in a variety of chemicals (herein agrochemicals). The barcode is readout by detecting melting peaks of organic nanoparticles using differential scanning calorimetry. This method has high labeling capacity due to small sizes of nanoparticles, sharp melting peaks, and large scan range of thermal analysis. The in-product barcode can be effectively used to protect agrochemical products from being counterfeited due to its large coding capacity, technical readiness, covertness, and robustness. Full article
(This article belongs to the Special Issue Engineered Nanomaterials in the Environment)
Show Figures

Figure 1

Figure 1
<p>In product barcodes based on phase change nanomaterials for agrochemical products.</p>
Full article ">Figure 2
<p>(<b>A</b>) Scanning electron microscope (SEM) and (<b>B</b>) transmission electron microscope (TEM) images of polymer encapsulated nanoparticles. (<b>C</b>) Fourier Transform infrared spectrometer (FTIR) spectra of paraffin (1) and polymer-encapsulated paraffin nanoparticles (2); size distribution of polymer encapsulated nanoparticles (C inset).</p>
Full article ">Figure 3
<p>Differential scanning calorimetry (DSC) curve of four paraffin nanoparticles in paclobutrazol (<b>A</b>); mass dependent peak area of paraffin nanoparticles (<b>B</b>); peak areas (<b>C</b>) and peak widths (<b>D</b>) of paraffin nanoparticles at different heating rates.</p>
Full article ">Figure 4
<p>Calculated phase diagram and DSC curves (inset). (<b>A</b>) palmitic acid and myristic acid; (<b>B</b>) the eutectic mixture of two acids (<span class="html-italic">i.e.</span>, palmitic acid and myristic acid) and stearic acid; (<b>C</b>) the eutectic mixture of three acids (<span class="html-italic">i.e.</span>, palmitic acid-myristic acid-stearic acid) and lauric acid.</p>
Full article ">Figure 5
<p>Barcoding agrochemical products with a library of barcodes formed by five types of organic solid nanoparticles.</p>
Full article ">
Back to TopTop