?w ii! A. than». «a»! .Q‘, «w. w} I I. l... t 1 3‘. «.1... . it}: 3.... z .. [‘42: 5!.- a. h? IIIFJUofihvé [MI 27 is}! Luz} . F1. 5.3.. u. o «avert... «9. A. 9.111.: 57.. .l. :61... 115(- !) ,qu .31. , 1.“. 3i :1 18.3! inn. Q. 0- " LIBRARY AXN Michigan State M University This is to certify that the dissertation entitled THE STUDY AND EXPLORATION ABOUT NANOTOXlCITY OF OXIDE NANOPARTICLES ON ESCHERICHIA COLI AND EXIGUOBACTERIUM SIBIRICUM presented by Yu Yang has been accepted towards fulfillment of the requirements for the Ph.D. degree in Crop and Soil Science - C 7 Mir Professbfs Sijnature 344 8 a £00? ’ / Date MSU is an Affinnative Action/Equal Opportunity Employer Iv-u-n-.—.-. PLACE IN RETURN Box to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier duedate if requested. DATE DUE DATE DUE DATE DUE ljf1.;\ Z 0 2012 (W <1 'I ‘ JuNOf 8‘ m3 ‘ 5/08 K:IProj/Aoc&Pres/ClRClDaIeDue.indd THE STUDY AND EXPLORATION ABOUT NANOTOXICITY OF OXIDE NANOPARTICLES ON ESCHERICHIA COLI AND EXIGUOBACTERIUM ' SIBIRICUM By Yu Yang A DISSERTATION Submitted to Michigan State University In partial fulfillments of requirements for the degree of DOCTOR OF PHILOSOPHY Crop and Soil Science 2009 ABSTRACT THE STUDY AND EXPLORATION ABOUT NANOTOXICITY OF OXIDE NANOPARTICLES ON ESCHERICHIA COLI AND EXIGUOBACTERIUM SIBIRICUM By Yu Yang Nanoparticles (NPS) have been used in a variety of areas, including cosmetics, pigments, water purification, environmental remediation and drug delivery. However, their toxicological and environmental effects have not been adequately investigated. I found that ZnO and TiOz NPS reduced cell viability of Escherichia coli K12 (Gram negative) and Exiguobacterium sibiricum strain 255-15 (Gram positive) in a dose-dependent manner, while SiOz and Fe3O4 NPS did not in a 12 h exposure. Among the test NPS, ZnO NPS were more toxic to E. coli than to E. sibiricum, while TiOz NPS were more toxic to E. sibiricum than to E. coli. The smallest ZnO or TiOz NPS, 5-10 mn, were more toxic than the larger particles to both bacteria. Zn2+ dissolved from ZnO NPS in both LB and TSB broth, reached saturation at 100-110 ppm. The release of Zn2+ ion played a more important role for ZnO NP toxicity to E. coli than E. sibiricum, due to the greater sensitivity of E. coli to Zn”, while both Zn2+ and ZnO NPS showed the toxicity to E. sibiricum. Si02 and TiOz NPS were internalized in E. sibiricum but not in E. coli cells as observed by transmission electron microscopy-dispersive spectroscopy, while ZnO NPS were found internal to both of them. No particle larger than 150 nm was found intracellularly. NPS that can be internalized by microbes may possess the stronger toxicity than those not internalized, which is consistent with my Observations. ZnO and TiOz NPS induced production of reactive oxygen species (ROS) in E. coli K12 and E. sibiricum in a dose-dependent manner. Both material-effect and size-effect were evidenced in ROS induction studies in that the capability for ROS induction increased generally from SiOz to TiOz to ZnO, and that the smaller TiOz NPS induced more ROS than the larger ones. More ROS was induced by SiOz NPS in E. sibiricum than in E. coli, while ZnO NPS induced more ROS in E. coli than in E. sibiricum, which is consistent with the viability results. Both TiOz NPS and ZnO NPs caused significant DNA damage in lymphocytes as measured by the comet assay, and TiOz NPS were more genotoxic than ZnO NPS. The latter is probably due to the Solubility or smaller size of the ZnO NPS. Using microarray technology, 1 found that SiOz, HQ», and ZnO NPS influenced gene expression of E. coli and E. sibiricum in a variety of function (COG) categories. Genes in the “transport” category were most affected by all treatments. When genes for amino acid and carbohydrate transport were repressed, genes for metal iron transport were induced. Other important categories with altered expression due to NPS were “carbohydrate metabolism”, “lipid metabolism”, “amino acid metabolism” and “DNA repair”. The latter may be caused by ROS damage to the membrane, amino acids in proteins, and DNA. NPs also led to up-regulation of stress-related genes and down-regulation of cellular respiration, also evidence for ROS induction. Copyright by Yu Yang 2009 To my parents Zhaolan Liang and Xiaoping Yang ACKNOWLEGEMENTS I would like to express my deepest gratitude to my advisors, Drs. James Tiedje and Syed Hashsham, for their continuous guidance, mentoring, and support through my Ph.D study. I learn a lot by their examples and I believe their infectious enthusiasm and passion about science will always inspire me in my future research. I so appreciate their generous encouragement and confidence, especially when I struggled with my research and almost lost confidence in myself. I can not imagine, without their constructive advice and comments, how I can finish this work. I would also like to extend my great appreciation to my other guidance committee members: Drs. Terence Marsh and Stephen Boyd, for their valuable comments and suggestions on this work. Thanks to all the members of Tiedje’s lab and Hashsham’s lab, for their invaluable discussions and kind help during the past five years, who make this period of time more colorful and give me so many precious memories. I would like to thank Drs. Alicia Pastor and Xudong Fan for their constant assistance and advice about transmission electron microscopy and sample preparation. I also want to say thanks to Dr. Alok K. Pandey for his valuable patience and assistance about comet assay and Mr. Yang-lyang Pan for his generous help about research supports. I can never thank my parents, Zhaolan Liang and Xiaoping Yang enough. Without their love, support, motivation and encouragement over these years, I would have never achieved my dream. NO matter what happens, they always trust me and have confidence in me. No matter when I feel weak inside and turn to them, they always open arm and vi give me their generous love. I hope I can make them proud of me. vii TABLE OF CONTENTS LIST OF TABLES ............................................................................................................................ x LIST OF FIGURES ......................................................................................................................... xi ABBREVIATIONS ........................................................................................................................ xii CHAPTER 1. ................................................................................................................................... I INTRODUCTION: TOXICITY OF NANOMATERIALS TO MICROORGANISMS Abstract ............................................................................................................................... 2 Introduction ......................................................................................................................... 3 Classes ofNMs and their toxicity ....................................................................................... 5 Mechanisms of NP toxicity ............................................................................................... 11 Conclusion ........................................................................................................................ 15 References ......................................................................................................................... 20 CHAPTER 2 .................................................................................................................................. 34 TOXICITY OF OXIDE NANOPARTICLES ON ESCHERICHIA COLI AND EXIGUOBACTERIUM SIBIRICUM Abstract ............................................................................................................................. 35 Introduction ....................................................................................................................... 36 Materials and methods ...................................................................................................... 37 Results ............................................................................................................................... 39 Discussion ......................................................................................................................... 44 References ......................................................................................................................... 59 CHAPTER 3. ................................................................................................................................. 65 INDUCTION OF INTRACELLULAR REACTIVE OXYGEN SPECIES BY NANOPARTICLES AND MEASUREMENT OF GENOTOXICITY BY COMET ASSAY Abstract ............................................................................................................................. 66 Introduction ....................................................................................................................... 67 Materials and methods ...................................................................................................... 68 Results and discussion ...................................................................................................... 71 References ......................................................................................................................... 82 CHAPTER 4. ................................................................................................................................. 89 INFLUENCES OF OXIDE NANOPARTICLES ON GENE EXPRESSION OF ESCHERICHIA COLIKIZ Abstract ............................................................................................................................. 90 Introduction ....................................................................................................................... 91 viii Materials and methods ...................................................................................................... 92 Results ............................................................................................................................... 98 Discussion ....................................................................................................................... I 13 References ....................................................................................................................... 120 CHAPTER 5 ................................................................................................................................ 133 COMPARATIVE ANALYSIS OF DIF F ERENTIALLY EXPRESSED GENES EXIGUOBACTERIUM SIBIRICUM FOLLOWING EXPOSURE TO OXIDE NPS Abstract ........................................................................................................................... I 34 Introduction ..................................................................................................................... 135 Materials and methods .................................................................................................... 136 Results ............................................................................................................................. 1 4 I Discussion ....................................................................................................................... 155 References ....................................................................................................................... I 63 ‘ix LIST OF TABLES Table 1.1 Examples of NMs showing antibacterial activity ............................................. 18 Table 2.1 The concentration of Zn2+ ion in LB or TSB medium from ZnO NPS ......... 53 Table 2.2 The toxicity of ZnClz to E. coli and E. sibiricum ............................................. 53 Table 3.1 Measurement of tail length, tail DNA and olive tail moment by TiOz NPS in comet assay ................................................................................................................................. 79 Table 3.2 Measurement of tail length, tail DNA and olive tail moment by ZnO NPS in comet assay ................................................................................................................................. 80 Table 4.1 primer sequence for target genes for validation .............................................. 118 LIST OF FIGURES Figure 1.1 Images of nanomaterials ....................................................................................... 17 Figure 2.1 The effect Of ZnO NPs on E. coli and E. sibiricum ........................................ 49 Figure 2.2 The effect of TiOz NPs on E. coli and E. sibiricum ........................................ 50 Figure 2.3 The toxicity of iron oxide NPs to E. coli and E. sibiricum ........................... 51 Figure 2.4 The influence Of SiOz NPs on E. coli and E. sibiricum ................................. 52 Figure 2.5 The toxicity of TiOz NP suspension and soluble fraction to E. coli and E. sibiricum ...................................................................................................................................... 54 Figure 2.6 The toxicity of ZnO NP suspension and soluble fraction to E. coli and E. sibiricum ...................................................................................................................................... 55 Figure 2.7 TEM and EDS images of SiOz NPs in E. sibiricum ....................................... 56 Figure 2.8 TEM and EDS images of TiOz NPs in E. sibiricum ....................................... 57 Figure 2.9 Internalization of ZnO NPs in E. coli and E. sibiricum ................................. 58 Figure 3.1 ROS induction by TiOz and ZnO NPs in E. coli and E sibiricum ............... 77 Figure 3.2 Human lymphocytes treated with TiOz NPs for 6 h exposure ...................... 81 Figure 4.1 The functional categories of the gene profiles in SiOz, TiOz and ZnO NP for E. coli ................................................................................................................................... 119 Figure 5.1 The upregulated and downregulated gene profiles in three different treatments .................................................................................................................................. 160 Figure 5.2 The functional categories of the gene profiles in SiOz, TiOz and ZnO NP for E. sibiricum ........................................................................................................................ 161 Figure 5.3 The functional categories of the shared gene profile among SiOz, TiOz and ZnO NP exposure .................................................................................................................... 162 xi Nanoparticles Nanomaterials Carbon nunotubes Quantum dots Reactive oxygen species Transmission electron microscopy Energy dispersive spectrometry Luria-Bertani Tryptic soy broth Tryptic soy agar 2’, 7’-dichlorofuorescin diacetate Polymerase chain reaction ATP-binding cassette Phosphotransferase system Coenzyme A Tricarboxylic acid ABBREVIATIONS NPs NMs CNTs QDs ROS TEM EDS LB TSB TSA DCFH-DA PCR ABC PTS COA TCA xii CHAPTER 1 INTRODUCTION: TOXICITY OF NAN OMATERIALS TO MICROORGANISMS Abstract The unique and advanced functions of nanomaterials (NMs) make them broadly useful and powerful for research, industrial development and many products. Their use, however, also brings uncertainties regarding their bio-safety and environmental impacts. The goal of this study is to introduce the main types of W3 and review their environmental adverse effect, with a particular emphasis on microorganisms. I also collate published papers regarding the antibacterial activity of different NMs, and describe several possible mechanisms of NM toxicity. Production of reactive oxygen species (ROS) is the most favored explanation for NP toxicity. ROS can attack DNA, amino acids in proteins and lipids in membranes, causing extensive damage to cells. Direct contact between NMs and microorganisms is also necessary for NP toxicity. Introduction Nanotechnology, a broad interdisciplinary field revolutionizing research development and new product and process, is one of the most rapidly growing fields. In nanotechnology and nanoscience, NMs, nanotools and nanodevices are the important segments, and the global nanotech market is estimated to grow to $25.2 billion by 2011 (75). The Royal Society stated that “Nanotechnologies are the design, characterization, production and application of structures, devices and systems by controlling shape and size at nanometer scale”, when there have also been a few definitions offered by other organizations (25). Most of these definitions refer only to the size range or new functionalities, but neglect risks and uncertainties which also need to be considered and resolved. The development of nanotechnology and nanoscience requires multidisciplinary integration of physics, chemistry, biology, and engineering, and mainly consists of nano-medicine, nanO-fabrication, nano-metrology and nano-materials (59). Manufactured NMs are defined by the US. National Nanotechnology Initiative as materials that have at least one dimension in the 1 to 100 nm range (88). Recently more and more NMs have been synthesized and applied in various fields, such as environmental remediation, drug delivery, cosmetics, electronics, energetics and photonics (3). For instance, 2954 kg of fullerenes were synthesized world-wide in 2003 (10). In 2005, more than ten billion dollars was spent internationally on nanotechnology development, and it is estimated that by 2015 the global market for nanotechnology and NMs will be around $1 trillion (35). Because of wide and abundant applications, NMs may enter environment during their life cycle and end up as environmental contaminants Of water, air and soil (93). The concerns about the toxicological and environmental impacts of NMs have been increasing rapidly, and the research conducted about safety of NMs remains focused on the issues related to occupational health and human safety. The concerns about environmental safety of NMs need attention; these materials are too small to detect and track in environment. The entrance of NMs to environment can be divided into intentional releases and unintentional releases, such as atmospheric emission and liquid waste streams. Nanoparticles (NPs), one of the most useful subsets of NMs, are widely applied in water purification, catalysis, pigments, food, cosmetics, and lubrication (3, 31, 38, 86, 116). NPs have natively existed in nature, while the exposure to the NPs may have some adverse effect on the environmental microorganisms(81). The mechanisms of NP toxicity are not yet clear, and the studies associated on their potential influence on public health and environments are inadequate. In this chapter, I focus on the toxicity Of NPS to microorganisms. Understanding NP toxicity to microorganism is important to evaluate the risk of NPs to environment, since microorganisms play a fundamental role of the food chain as the lowest level and are responsive for much of biogeochemical cycles. Moreover, they grow rapidly and are sensitive to some toxicants, which make them as a good test model for NP toxicity. These studies will supply helpful information for regulatory policy concerning the manufacture, application and disposition of NPs. Classes of NMs and their toxicity NMs have always existed in‘the environment, from both natural and anthropogenic sources. The natural NMs include clays and particulate organic matter in soil, volcanic ash and forest-fire smoke, and, colloid and ocean spray in aquatic systems. The engineered NMs possess many unique characteristics, which are attributable to their small size, large surface area. chemical composition, surface structure, solubility, shape, and aggregation. The main types Of manufactured NMS include: Fullerene The manufacture of carbon-base fullerenes, spherical molecules comprising 60 carbon atoms, opened the gate of the nanotechnology era (58). Its configuration is like a football with 20 hexagons and 12 pentagons and its diameter is about 1 nm (Figure l.1(a)). Underivatized fullerenes are extremely insoluble in water, and subject to precipitation and aggregation (4, 54). Fullerene has been applied in polymer and structural composites, electromagnetic shielding, electron field emitters, super capacitors and batteries, due to their high conductivity, high surface area and unique electronic properties (3, 74). F ullerenes are redox active and can generate ROS, which may cause damage to DNA, membranes and amino acids (20, 100). Fullerenes have been observed to alter the membrane lipid composition, phase transition temperature and membrane fluidity in Pseudomonas putida and Bacillus subtilis (100). The antibacterial activity of fullerene on Escherichia coli and B. subtilis has been reported in literature, as a decreased aerobic respiration rate (28, 63). Surface-modified fullerene can generate singlet oxygen, causing lipid peroxidation and exhibiting antibacterial activities (48, 64, 67). Carboxyfullerene caused the inhibition on Gram positive species by damage to cell membranes, i. e. Staphylococcus spp., Streptococcus spp., Enterococcus faecalis, but not for Gram negative E. coli, Pseudomonas aeroginosa, Salmonella typhmuriumi and Klebsiella pneumoniae (108, 109). However, toxicity of a carboxy-fullerene derivative to E. coli was detected in a dose-dependent manner (107). Some other fullerene derivatives were also studied for their antibacterial activities. For instance, one positively charged fullerene derivative inhibited the growth of Mycobacterium tuberculosis, and alkylated fullerene derivatives could inhibit the growth of E. faecalis, S. aureus, S. epidermidis, and Enterococcus hirae (11, 67). Their toxicity to E. coli was explained as respiratory chain inhibition and hydrogen peroxide production (68, 69). Moreover, fullerene derivatives had a mutagenic effect on S. typhimurium (6). It was also reported that fullerene can alter the microbial community in soil (106). Nanotubes and nanowires Carbon nanotubes (CNTS), first observed by Sumio Iijima in 1991, are a particular form of fullerenes, including single-walled carbon nanotubes (SWCNTS) and multi-walled carbon nanotubes (MWCNTS) (Figure 1.1(b)) (44). Both SWCNBs and MWCNTS show antibacterial activity: direct contact between SWCNBs and bacteria can cause membrane damage and lead to cell death (49, 85). Dr Elimelech’s group published a serial of papers about CNT toxicity. They provided the direct evidence that both SWCNTS and SWCNTS exhibited strong antimicrobial activity on E. coli, and SWCNBs were more toxic to bacteria than MWCNTs (50, 51). They also reported that these CNBS could regulate gene expression of E. coli and induce stress-related genes, and that the size of CNT could influence the antibacterial effects (49). Meanwhile, the surface coating is also important for CNT antimicrobial activity (85). Nanowires are ultrafine wires or linear arrays of dots formed by self assembly, and used as interconnectors in nanoelectronic devices (25). There are not many studies of nanowire toxicity, and perhaps they are less likely to be toxic due to their long dimension. Nanoparticles One important subset of NMs are NPs, which are particles with a diameter less than 100 nm. The materials of NPs include metal, metal oxides and silica. The metal oxide NPs can be consisted of individual oxides, such as ZnO, T102, CeOz and CrOz, and binary oxides, such as BaTi03 and LiCOOz. NPS of Silicon oxide (SiOz), Titanium oxide (TiOz) and Zinc oxide (ZnO) are the popular ones among the oxide NPs. SiOz is used as an additive in food, drugs, cosmetics and printer toners (Figure l.1(a)). It is also a good semiconductor and an electrical insulator. TiOz is used as a photocatalyst for disinfection of bacteria, and a number of the papers explored this subject (9, 14, 66, 72, 94, 96, 104). It is also employed as a pigment in paints and a sunscreen product in cosmetics. ZnO is widely used in pigments, semiconductors, and cosmetics. It was estimated that globally 1000 tons of NPs were used in sunscreen products in 2004 (10). There have been a ntunber of reports on the toxicity Of SiOz, TiOz and ZnO NPS. SiOz NPs are the least toxic among these three NPs. Their antibacterial activity on E. coli and B. subtilis was reported while some studies indicated that they were not toxic (1). Due to the wide application of TiOz NPs, their toxicity on microorganisms has been well studied. The TiOz NP photocatalytic inactivation of E. coli, S. aureus and E. faecalis was reported, and scientists also demonstrated antibacterial activity to Streptococcus spp., Enterococcus spp., Vibrio fischeri, Serratia marcescens, Bacillus megaterium and some fungi (29, 36, 76, 94, 95, 118). N-doped TiOz NPs and Ag-coated TiOz NPs also showed inhibition of E. coli and Micrococcus lylae, respectively (60, 61, 123). Matsunaga et al. reported that TiOz can oxidize COA and cause cell death (72, 73). Moreover, the cell membrane may be oxidized prior to the oxidation of COA, and a partial decomposition of outer membrane is followed by disordering of the cytoplasmic membrane, leading to cell death (19, 76, 104). Later, Saito et al. found rapid leakage of K+ ion and slow release of proteins and RNA. They explained that the cell death was due to a significant disorder in cell permeability (7, 96). The toxicity of Ti02 NPs can be influenced by NP size and light intensity (2, 39). ZnO NPs are probably the most toxic among these three NPs. They inhibited the growth of E. coli, S. aureus, V. fishceri, B. subtilis, S. typhimurium and Streptococcus agalactiae (43, 83, 92, 122). The release Of Zn2+ ion from ZnO NPs, direct contact between ZnO NPs and membranes, and ROS induction all contributed to their antibacterial activity. A NP size-effect was also demonstrated in ZnO toxicity studies, and internalization of ZnO was found in E. coli, S. agalactiae, Exiguobacterium sibiricum and S. aureus (Figurel.1 (b)) (12, 43, 120). Recently there have been a few reports about the antimicrobial activity of some other NPS, and on the interaction between NPS and microorganisms (12, 15, 29, 89, 90, 103). CuO NPs inhibited the growth of V. fischeri, and the growth of E. coli was found to be inhibited by N-doped ZrOz, CeOz, Ag/A1203 and Fe203 NPS (5, 15, 80, 105). Both CaO and MgO NPs demonstrated antibacterial activity to B. subtilis, E. coli, S. aureus, S. typhimurium and B. megaterium (41, 42, 65, 97, 98). Silver NPs are one of the most promising metal NPs for future applications. They are antibacterial, antimicrobial and antifungal (24). Silver NPS can also interact with DNA, and interrupt RNA replication and microbial reproduction (79, 110).The toxicity of Ag NPS to E. coli was reported by Sondi and Salopek-Sondi, and Ag NPS also inhibited the growth of Salmonella typhi, P. aeruginosa, and Vibrio cholerae (79, 90, 102). The toxicity of gold NPs was demonstrated to depend on surface coating. Cationic gold NPs are moderately toxic and anionic gold NPs are quite nontoxic (30, 117). Organic compound NPs also possess antibacterial activity, such as the growth inhibition of K. pneumonia, S. aureus, and Streptococcus mutans by polyethylenimine NPs and Ag-loaded polystyrene NPs (8, 89). Quantum dots (QDs). Quantum dots, which are small assemblies of metal, metal oxide or semiconductor materials in size of 2-10 nm, were first created in the early 19803 (25). Quite a few QDs contain some noble or transition metals, such as CdSe, InAs and ZnS. The components for QD shell or core can be quite different (18, 52). The application of QDs covers biomedical imaging, drug delivery and electronics industry (34). They can be coated by some organic molecules, which can influence their characteristics and toxicity (40). QDs can induce ROS and nick supercoiled DNA to cause genotoxicity (32, 46). Other NMs There are some other types of NMS, such as dendrimers and biopolymers. Few toxicity studies have been conducted on these NMS, perhaps due to their limiting applications. 10 Mechanisms of NP toxicity There have been a few hypothesized mechanisms Of NP toxicity. such as disruption of cell membranes, induction of ROS, dissolution of hazardous constituents and interruption Of energy production (54. 84). NPs may cause toxicity to bacteria via one or more Of these mechanisms. ROS induction Induction of ROS by NPs is currently the most favored explanation for the toxicity of NPs. In 1997, Rahman et al. found the increasing induction of ROS by nano-sized Ti02 in human and rat alveolar macrophages (91). That Ti02 NPs were toxic to some cells and caused oxidative stress in different cell types, including bacteria and mammal cells (2, 33, 62,101, 113-115). The chemical reactions by which Ti02 NPs induce ROS are showed below (16, 17, 123). In the first equation, a hole (e')/electron (h+) in the valence band and conduction band is induced by light (k<390 nm) and Ti02 NPs. The e' is absorbed onto the surface of a TiO2 NP and available for electron transfer, which can reduce 02 to 02' (reaction 2), and produce H202 by further reducing 02' (reaction 3). The OH radicals can be produced by the reaction of superoxide with H202 (reaction 4) and reduction of H202 by e' (reaction 5). The OH radicals can also be formed by the reactions of H20 or OH" with h+ and electron abstraction from absorbed oxidizable species by h+ (reaction 6 and 7). H202 11 can be produced by either a reductive pathway (reaction 3) or an oxidative pathway (reaction 8), while it can be decomposed into OH radicals by light (reaction 9). TiO2 + hv ——> TiO2 (e‘ + h*) (1) 6" + 02 —> 02' (2) 02' + 2H“ + e' —+ H202 (3) H202 + 02' —> °0H + OH' + 02 (4) e' + H202 —->°0H + OH' (5) h” + OH' ——>°OH (6) h+ + H2O —+ °OH + IT“ ’ (7) 2h+ + 2H2O —+ 2H“ + H202 (8) hv + H202 —» 2°0H (9) An interfacial charge transfer has been reported to occur between Ti02 NPs and E. coli membrane (82). Reaction 10 and 11 exhibit the surface charge trapping on Ti02 NPs. During the process of interfacial charge transfer, h+ reacts with peptidoglycan (PGN) and lipo-polysaccharides (LPS) on microbial cell wall membranes, such as E. coli (reaction 12). Meanwhile, h+ can be scavenged by PE, LPS or E. coli during surface oxidation (reaction 13). e' + Ti(IV)0-H —+ Ti(III)0-H' (10) h+ +Ti(IV)0-H -2 Ti (Iv)O--H+ (11) hI+RH—>R°+H+ (12) Ti(1v)O--H++RH—»OH+R- +H+ (13) 12 ROS can be transferred through membranes, and oxidize fatty acids of phospholipids and cause lipid peroxidation. This process can change membrane permeability and fluidity, and make bacteria more vulnerable to osmotic stress and interrupt nutrient uptake (13). ROS can also oxidize and damage DNA to cause genotoxicity. For instance, there have been a few reports about the DNA lesions caused by ROS induction by Ti02 NPs (22, 33, 37, 57, 70, 111-113). Moreover. ROS can oxidize amino acids in proteins, and especially attack sulfur-iron assembly, which can release Fe ions, induce Fenton chemistry and lead to more production of ROS (45). The induction Of ROS by ZnO NPs was also hypothesized as an explanation for their toxicity, although the mechanisms for ZnO NP toxicity are not yet well understood (99). ZnO NP toxicity may be caused by the binding of the NPs to the bacterial surface due to electronic forces (103, 122). The internalization of NPs is not required for ROS induction. ROS can be transferred into bacteria via direct contact between NPs and their membranes, although the lifetime for ROS is very short (10'9 second). Release of toxic components from NPs Hazardous components can be released by some NMs. Silver NPs are a good example, as release of silver ions has been hypothesized as one of the mechanisms for toxicity of silver NPs (79). That the interaction between silver ions and thiol groups in proteins can inactivate some enzymes (71). Moreover, silver ions can also attack membranes and prevent DNA replication (27). ZnO NPs can release Zn2+ ion, which is 13 antibacterial (121). Kirchner et al. reported that QDs (CdSe/ZnS) released toxic Cd”, which increased their toxicity (53). Meanwhile, both free Cd and Se from QDS was found inside S. aureus before QDs was internalized, which confirmed the release of components from QDS (56). Damage to membrane integrity Membrane integrity is important for microorganisms to maintain cell activity, such as transport and energy transduction, and to protect cells from potential damage. Some NMs can attack bacterial membranes directly. Zn0 NPs were internalized into E. coli cells (12). E. coli and E. sibiricum cells with broken membrane and leaking cytoplasm were found when they were exposed to ZnO NPs in our studies. QDs were also found to enter bacterial cells (55). However, there is no clear explanation for how NP internalization occurs, since NPs are still too big for channels or transporters to transport through membranes. Moreover, Ag NPs can also interact with sulphur-containing amino acids in membranes, prohibiting transport through membranes and causing cell death (102, 119). The embedding of silicon NPs and fullerene in membranes was also reported (47, 107). ROS can disrupt membranes too, via oxidizing lipids or proteins in membrane. Cell membrane damage by direct contact of CNTs is proposed as the main mechanism for CNT cytotoxicity. Damage to nucleic acid 14 Some NMs have characteristics of interacting with DNA, and they can be used to label or cleave DNA. For example, QDs have been used for imaging applications via tagging DNA, and they were also found to nick DNA (21, 23, 32). ROS induced by NPs can damage DNA, such as by Ti02 and ZnO NPs (22). Conclusion The NM toxicity studies require multidisciplinary collaboration of physicist, chemist, microbiologist and other experts. On one hand, when microbiologists focus on antibacterial activity of NMS, they may neglect the characterization of NPs. However, it is essential to characterize the physical or physieochemieal properties of NMs before any toxicity studies, since their features are likely critical to any toxicity. Therefore, the measurement Of the NP properties should be the first step for studying their toxicity. The main characteristic Of NPs is their size. which may extend down to individual atoms or molecules, and much smaller than cells. The small size can influence other physicochemical properties of materials and increases the Opportunity of uptake and interaction with or within bacteria. Surface area is another important factor in toxic effects. Normally when the size decreases, the surface area increases and there are more reactive groups. The surface group is dependent on the chemical composition and surface structure of NPS, such as electronic properties and coating. which determines the solubility of NPS. hydrophobic or hydrophilic. lipophobic or lipophilic, or active or 15 passive. The other main properties of NPs that may affect their toxicity include agglomeration state, shape, crystal structure, chemical composition, surface chemistry, surface charge and porosity (88). Standard and characterized NPS are critical for NP toxicity studies, as well are standardized and validated test methods. 0n the other hand, physicists and chemists should use more relevant biological methods to explore toxicity study at molecular level. Simple viability tests and growth inhibition are not very helpful to answer the questions about mechanisms of NP toxicity. Use of genetics and proteomics can supply more details about bacterial reaction to NPs, and reveal increasing complex that can be useful in refining experiments for exploration of NP fate, bioavailability and effect. There are still many unexplored areas for NP toxicity studies. For instance, there are few anaerobic studies, although it was found that fullerenes have no significant effect on anaerobic microbial communities (87). How microbial communities react to NMs still remains unknown, and it is very difficult to estimate the consequences of wide environmental impact during the exposure to NMs. Finally, it is critical to improve our understanding of the interaction between NPs and microorganisms. 16 w . ., .v. .2 . . u v .mw . «w... I. a Figure 1.] images of NPs. (a) TEM image of a fullerene (28).(b) TEM image of a SWCMT (49). (C) TEM image of 8102 NPs. (b) TEM image of internalized ZnO NPs in E. sibiricum. 17 Table 1.1 Examples of NMs showing antibacterial activity NMs Microorganism Reference Fullerenes E. coli, B. subtilis, and P. putida (94), (64), (63). (28), (26), fullerene E. coli, S. typhimurium, and M. tubercolosis (6), (11), (68), (69) derivatives carboxyfullerene E. coli, S. pyogenes, Staphylococcus spp., (107), (109), (108) Streptococcus spp., E. faecalis, P. aeruginosa, S. typhi, and K. pneumoniae SW and MWNTs E. coli, M lysodeikticus, and S. aureus (85), (49), (51), (50) Carbon Nanohom E. coli (77) Carbon NPs K. pneumoniae (78) QDS E. coli and B. subtilis (55) Gold NPs E. coli (30), (117) CuO NPS V. fischeri (80), (36) Si02 E. coli and B. subilis (I) TiO2 E. coli, M. luteus, B. subtilis, Streptococci spp., S. (95), (94), (9), (116), marcescens, Enterococcus spp., Saccharomyces (76), (118), (29), Serevisiae, S. aureus, E. faecalis, and V. fischeri (104), (l), (16), (7), (96), (72), (39), (124), (36) N-doped Ti02 E. coli (60), (61) A g-coated TiO2 M lylae (123) Silver E. coli, P. aeruginosa, V. cholera, and s. typhi (102), (79), (90) N-doped Zr02 E. coli (61) ZnO S. agalactiae, S. aureus, V. fischeri, B. subilis, and (98), (92), (l), (122), S. typhimurium (36), (97), (83), (120), (43), (80) MgO B. subtilis, E. coli, S. aureus, B. megaterium, (98), (42), (41), (103), and S. typhimurium (97), (65) CaO B. subtilis, E. coli, S. typhimurium, and S. aureus (98), (97) Ce02 E. coli (105) Ag/Al203 E. coli (15) Fe203 E. coli (5) Silica, silica/iron E. coli (117) oxide Polyethylenimine Streptococcus mutans (8) NPs 18 Table 1.1 cont Ag-loaded polystyrene NPs K. pneumonia and S. aureus (89) 19 References 1. Adams, L. K., D. Y. Lyon, and P. J. J. Alvarez. 2006. Comparative coo-toxicity of nanoscale Ti02, Si02, and ZnO water suspensions. Water Research 40:3527-3532. 2. Adams, L. K., D. Y. Lyon, A. McIntosh, and P. J. J. Alvarez. 2006. Comparative toxicity of nano-scale Ti02, Si02 and ZnO water suspensions. Water Science and Technology 54:327-334. 3. Aitken, R. J., M. Q. Chaudhry, A. B. A. Boxall, and M. Hull. 2006. Manufacture and use of nanomaterials: current status in the UK and global trends. Occupational Medicine-Oxford 56:300-306. 4. Andrievsky, G. V., V. K. Klochkov, E. L. Karyakina, and N. 0. Mchedlov-Petrossyan. 1999. Studies of aqueous colloidal solutions of fullerene C-60 by electron microscopy. Chemical Physics Letters 300:392-396. 5. Auffan, M., W. Achouak, J. Rose, M. A. Roncato, C. Chaneac, D. T. Waite, A. Masion, J. C. Woicik, M. R. Wiesner, and J. Y. Bottero. 2008. Relation between the redox state of iron-based nanoparticles and their cytotoxicity toward Escherichia coli. Environmental Science & Technology 42:6730-6735. 6. Babynin, E. V., I. A. Nuretdinov, V. P. Gubskaya, and B. I. Barabanshchikov. 2002. Study of mutagenic activity of fullerene and some of its derivatives using his plus reversions of Salmonella typhimuriurn as an example. Russian Journal of Genetics 38:359-363. 7. Bekbolet, M., and C. V. Araz. 1996. Inactivation of Escherichia coli by photocatalytic oxidation. Chemosphere 32:959-965. 8. Beyth, N., I. Yudovin-Farber, R. Bahir, A. J. Dumb, and E. Weissa. 2006. Antibacterial activity of dental composites containing quaternary ammonium polyethylenimine nanoparticles against Streptococcus mutans. Biomaterials 27:3995-4002. 20 10. 11. 12. 13. 14. 15. 16. 17. 18. Block, S. S., V. P. Seng, and D. W. Goswami. 1997. Chemically enhanced sunlight for killing bacteria. Journal of Solar Energy Engineering-Transactions of the Asme 119:85-91. Borm, P. J. A., D. Robbins, S. Haubold, T. Kuhlbusch, H. Fissan, K. Donaldson, R. Schins, V. Stone, W. Kreyling, J. Lademann, J. Krutmann, D. Warheit, and E. Oberdorster. 2006. The potential risks of nanomaterials: a review carried out for ECETOC. Paricle and Fibre Toxicology 3. Bosi, S., T. Da Ros, S. Castellano, E. Banfi, and M. Prato. 2000. Antimycobacterial activity of ionic fiillerene derivatives. Bioorganic & Medicinal Chemistry Letters 10: 1043-1045. Brayner, R., R. Ferrari-Iliou, N. Brivois, S. Djediat, M. F. Benedetti, and F. Fievet. 2006. Toxicological impact studies based on Escherichia coli bacteria in ultrafine ZnO nanoparticles colloidal medium. Nano Letters 6:866-870. Cabiscol, E., J. Tamarit, and J. Res. 2000. Oxidative stress in bacteria and protein damage by reactive oxygen species. International Microbiology 3:3-8. Cai, R., K. Hashimoto, K. Itoh, Y. Kubota, and A. Fujishima. 1991. Photokilling of malignant-cells with ultrafine Ti02 powder. Bulletin of the Chemical Society of Japan 64:1268-1273. Chang, Q. Y., L. Z. Yan, M. X. Chen, H. He, and J. H. Qu. 2007. Bactericidal mechanism of Ag/Al203 against Escherichia coli. Langmuir 23:11197-11199. Cho, M., H. Chung, W. Choi, and J. Yoon. 2004. Linear correlation between inactivation of E-coli and 0H radical concentration in Ti02 photocatalytic disinfection. Water Research 38: 1069-1077. Cho, M., H. M. Chung, W. Y. Choi, and J. Y. Yoon. 2005. Different inactivation behaviors of MS-2 phage and Escherichia coli in Ti02 photocatalytic disinfection. Applied and Environmental Microbiology 71:270-275. Dabbousi, B. 0., J. RodriguezViejo, F. V. Mikulec, J. R. Heine, H. Mattoussi, 21 19. 20. 21. 22. 23. 24. 25. 26. R. Ober, K. F. Jensen, and M. G. Bawendi. 1997. (CdSe)ZnS core-shell quantum dots: Synthesis and characterization of a size series of highly luminescent nanocrystallites. Journal of Physical Chemistry B 101:9463-9475. Dadjour, M. F., C. Ogino, S. Matsumura, and N. Shimizu. 2005. Kinetics of disinfection of Escherichia coli by catalytic ultrasonic irradiation with TiOz. Biochemical Engineering Journal 25:243-248. Dhawan, A., J. S. Taurozzi, A. K. Pandey, W. Q. Shan, S. M. Miller, S. A. Hashsham, and V. V. Tarabara. 2006. Stable colloidal dispersions of C-60 fullerenes in water: Evidence for genotoxicity. Environmental Science & Technology 40:7394-7401 . Dubertret, B., P. Skourides, D. J. Norris, V. Noireaux, A. H. Brivanlou, and A. Libchaber. 2002. In vivo imaging of quantum dots encapsulated in phospholipid micelles. Science 298: 1759-1762. Dunford, R., A. Salinaro, L. Z. Cai, N. Serpone, S. Horikoshi, H. Hidaka, and J. Knowland. 1997. Chemical oxidation and DNA damage catalysed by inorganic sunscreen ingredients. Febs Letters 418:87-90. Dyadyusha, L., H. Yin, S. Jaiswal, T. Brown, J. J. Baumberg, F. P. Booy, and T. Melvin. 2005. Quenching of CdSe quantum dot emission, a new approach for biosensing. Chemical Communications:3201-3203. Elechiguerra, J. L., J. L. Burt, J. R. Morones, A. Camacho-Broagado, ‘X. Gao, H. H. Lara, and M. J. Yacaman. 2005. Interaction of silver nanoparticles with HIV-1. Journal of Nanobiotechnology 3. Engineering, T. R. S. a. t. R. A. o. 2004. Nanoscience and nanotechnologies: opportunities and uncertainties. Fang, J. S., D. Y. Lyon, M. R. Wiesner, J. P. Dong, and P. J. J. Alvarez. 2007. Effect of a fullerene water suspension on bacterial phospholipids and membrane phase behavior. Environmental Science & Technology 41:2636-2642. 22 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. Feng, Q. L., J. Wu, G Q. Chen, F. Z. Cui, T. N. Kim, and J. 0. Kim. 2000. A mechanistic study of the antibacterial effect of silver ions on Escherichia coli and Staphylococcus aureus. Journal of Biomedical Materials Research 52:662-668. Fortner, J. D., D. Y. Lyon, C. M. Sayes, A. M. Boyd, J. C. Falkner, E. M. Hotze, L. B. Alemany, Y. J. Tao, W. Guo, K. D. Ausman, V. L. Calvin, and J. B. Hughes. 2005. C-60 in water: Nanocrystal formation and microbial response. Environmental Science & Technology 39:4307-4316. Fu, G. F., P. S. Vary, and C. T. Lin. 2005. Anatase Ti02 nanocomposites for antimicrobial coatings. Journal of Physical Chemistry B 109:8889-8898. Goodman, C. M., C. D. McCusker, T. Yilmaz, and V. M. Rotello. 2004. Toxicity of gold nanoparticles functionalized with cationic and anionic side chains. Bioconjugate Chemistry 15:897-900. Goswami, D. Y., D. M. Trivedi, and S. S. Block. 1997. Photocatalytic disinfection of indoor air. Journal of Solar Energy Engineering-Transactions of the Asme 119:92-96. Green, M., and E. Howman. 2005. Semiconductor quantum dots and free radical induced DNA nicking. Chemical Communications: 121-123. Gurr, J. R., A. S. S. Wang, C. H. Chen, and K. Y. Jan. 2005. Ultrafine titanium dioxide particles in the absence of photoactivation can induce oxidative damage to human bronchial epithelial cells. Toxicology 213:66-73. Hardman, R. 2006. A toxicologic review of quantum dots: Toxicity depends on physicochemical and environmental factors. Enviromnental Health Perspectives 114:165-172. Harrison, P. 2007. Emerging challenges: nanotechnology and the environment. GEO Year book 2007:61-68. Heinlaan, M., A. Ivask, I. Blinova, H. C. Dubourguier, and A. Kahru. 2008. Toxicity of nanosized andbulk ZnO, CuO and Ti02 to bacteria Vibrio fischeri 23 37. 38. 39. 40. 41. 42. 43. 44. 45. and crustaceans Daphnia magna and Thamnocephalus platyurus. Chemosphere 71:1308-1316. Hirakawa, K., M. Mori, M. Yoshida, S. Oikawa, and S. Kawanishi. 2004. Photo-irradiated titanium dioxide catalyzes site specific DNA damage via generation of hydrogen peroxide. Free Radical Research 38:439-447. Hoet, P. H. M., I. Bruke-Hohlfeld, and O. V. Salata. 2004. Nanoparticles - known and unknown health risks. Journal of Nanobiotechnology 2:12. Horie, Y., D. A. David, M. Taya, and S. Tone. 1996. Effects of light intensity and titanium dioxide concentration on photocatalytic sterilization rates of microbial cells. Industrial & Engineering Chemistry Research 35:3920-3926. Hoshino, A., K. Fujioka, T. Oku, M. Suga, Y. F. Sasaki, T. Ohta, M. Yasuhara, K. Suzuki, and K. Yamamoto. 2004. Physicochemical properties and cellular toxicity of nanocrystal quantum dots depend on their surface modification. Nano Letters 4:2163-2169. Huang, L., D. Li, Y. Lin, D. G. Evans, and X. Duan. 2005. Influence of nano-MgO particle size on bacterial acion against Bacillus subtilis var. nider. Chinses Science Bulletin 50:514-519. Huang, L., D. Q. Li, Y. J. Lin, M. Wei, D. G Evans, and X. Duan. 2005. Controllable preparation of nano-MgO and investigation of its bactericidal properties. Journal of Inorganic Biochemistry 99:986-993. Huang, Z. B., X. Zheng, D. H. Yan, G. F. Yin, X. M. Liao, Y. Q. Kang, Y. D. Yao, D. Huang, and B. Q. Hao. 2008. Toxicological effect of ZnO nanoparticles based on bacteria. Langmuir 24:4140-4144. Iijima, S. 1991. Helical microtubules of graphitic carbon. Nature 354:56-58. Imlay, J. A. 2003. Pathways of oxidative damage. Annual Review of Microbiology 57:395-418. 24 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 'Ipe, B. I., M. Lehnig, and C. M. Niemeyer. 2005. On the generation of free radical species from quantum dots. Small 1:706-709. Jang, H., L. E. Pell, B. A. Korgel, and D. S. English. 2003. Photoluminescence quenching of silicon nanoparticles in phospholipid vesicle bilayers. Journal of Photochemistry and Photobiology a-Chemistry 158:111-117. Kai, Y., Y. Komazawa, A. Miyajima, N. Miyata, and Y. Yamakoshi. 2003. [60]Fullerene as a novel photoinduced antibiotic. Fullerenes Nanotubes and Carbon Nanostructures 11:79-87. Kang, S., M. Herzberg, D. F. Rodrigues, and M. Elimelech. 2008. Antibacterial effects of carbon nanotubes: Size does matter. Langmuir 24:6409-6413. Kang, S., M. S. Mauter, and M. Elimelech. 2008. Physicochemical determinants of multiwalled carbon nanotube bacterial cytotoxicity. Environmental Science & Technology 42:7528-7534. Kang, S., M. Pinault, L. D. Pfefferle, and M. Elimelech. 2007. Single-walled carbon nanotubes exhibit strong antimicrobial activity. Langmuir 23:8670-8673. Kim, S., B. Fisher, H. J. Eisler, and M. G. Bawendi. 2004. Type-II Quantum Dots: CdTe/CdSe(Core/Shell) and CdSe/ZnTe(Core/Shell) Heterostructures. The Journal of American Chemical Society 125: 1 1466-11467. Kirchner, C., T. Liedl, S. Kudera, T. Pellegrino, A. M. Javier, H. E. Gaub, S. Stolzle, N. Fertig, and W. J. Parak. 2005. Cytotoxicity of colloidal CdSe and CdSe/ZnS nanoparticles. Nano Letters 5:331-338. Klaine, S. J., P. J. J. Alvarez, G. E. Batley, T. F. F ernandes, R. D. Handy, D. Y. Lyon, S. Mahendra, M. J. McLaughlin, and J. R. Lead. 2008. Nanomaterials in the environment: Behavior, fate, bioavailability, and effects. Environmental Toxicology and Chemistry 27 :1825-185 1. Kloepfer, J. A., R. E. Mielke, and J. L. Nadeau. 2005. Uptake of CdSe and CdSe/ZnS quantum dots into bacteria via purine-dependent mechanisms. Applied 25 56. 57. 58. 59. 60. 61. 62. 63. 64. and Environmental Microbiology 71:2548-2557. Kloepfer, J. A., R. E. Mielke, M. S. Wong, K. H. Nealson, G Stucky, and J. L. Nadeau. 2003. Quantum dots as strain- and metabolism-specific microbiological labels. Applied and Environmental Microbiology 69:4205-4213. Kong, S. J., B. M. Kim, Y. J. Lee, and H. W. Chung. 2008. Titanium dioxide nanoparticles trigger p53-mediated damage response in peripheral blood lymphocytes. Environmental and Molecular Mutagenesis 49:399-405. Kroto, H. W., A. W. Allaf, and S. P. Balm. 1991. C60 - Buckminsterfullerene. Chemical Reviews 91: 121 3-1235. Lehn, J. M. 2002. Toward self-organization and complex matter. Science 295:2400-2403. Li, Q., R. C. Xie, Y. W. Li, E. A. Mintz, and J. K. Shang. 2007. Enhanced visible-light-induced photocatalytic disinfection of E-coIi by carbon-sensitized nitrogen-doped titanium oxide. Environmental Science & Technology 41:5050-5056. Liu, Y., J. Li, X. F. Qiu, and C. Burda. 2007. Bactericidal activity of nitrogen-doped metal oxide nanocatalysts and the influence of bacterial extracellular polymeric substances (EPS). Journal of Photochemistry and Photobiology a-Chemistry 190:94-100. Long, T. C., J. Tajuba, P. Sama, N. Saleh, C. Swartz, J. Parker, S. Hester, G. V. Lowry, and B. Veronesi. 2007. Nanosize titanium dioxide stimulates reactive oxygen species in brain microglia and damages neurons in vitro. Environmental Health Perspectives 115: 1631-1637. Lyon, D. Y., L. K. Adams, J. C. Falkner, and P. J. J. Alvarez. 2006. Antibacterial activity of fullerene water suspensions: Effects of preparation method and particle size. Environmental Science & Technology 40:4360-4366. Lyon, D. Y., J. D. Fortner, C. M. Sayes, V. L. Colvin, and J. B. Hughes. 2005. 26 65. 66. 67. 68. 69. 70. 71. 72. Bacterial cell association and antimicrobial activity of a C-60 water suspension. Environmental Toxicology and Chemistry 24:2757-2762. Makhluf, S., R. Dror, Y. Nitzan, Y. Abramovich, R. Jelinek, and A. Gedanken. 2005. Microwave-assisted synthesis of nanocrystalline MgO and its use as a bacteriocide. Advanced Functional Materials 15:1708-1715. Maness, P. C., S. Smolinski, D. M. Blake, Z. Huang, E. J. Wolfrum, and W. A. Jacoby. 1999. Bactericidal activity of photocatalytic Ti02 reaction: Toward an understanding of its killing mechanism. Applied and Environmental Microbiology 65:4094-4098. Mashino, T., D. Nishikawa, K. Takahashi, N. Usui, T. Yamori, M. Seki, T. Endo, and M. Mochizuki. 2003. Antibacterial and antiproliferative activity of cationic fullerene derivatives. Bioorganic & Medicinal Chemistry Letters 13:4395-4397. Mashino, T., K. Okuda, T. Hirota, M. Hirobe, T. Nagano, and M. Mochizuki. 1999. Inhibition of E-coli grth by fullerene derivatives and inhibition mechanism. Bioorganic & Medicinal Chemistry Letters 9:2959-2962. Mashino, T., N. Usui, K. Okuda, T. Hirota, and M. Mochizuki. 2003. Respiratory chain inhibition by fullerene derivatives: Hydrogen peroxide production caused by fullerene derivatives and a respiratory chain system. Bioorganic & Medicinal Chemistry 11:1433-1438. Mates, J. M., and F. M. Sanchez-Jimenez. 2000. Role of reactive oxygen species in apoptosis: implications for cancer therapy. International Journal of Biochemistry & Cell Biology 32:157-170. Matsumura, Y., K. Yoshikata, S. Kunisaki, and T. Tsuchido. 2003. Mode of bactericidal action of silver zeolite and its comparison with that of silver nitrate. Applied and Environmental Microbiology 69:4278-4281. Matsunaga, T., R. Tomoda, T. Nakajima, N. Nakamura, and T. Komine. 1988. Continuous-sterilization system that uses photosemiconductor powders. Applied and Environmental Microbiology 54: 1330-1333. 27 73. 74. 75. 76. 77. 78. 79. 80. 81. Matsunaga, T., R. Tomoda, T. Nakajima, and H. Wake. 1985. Photoelectrochemical sterilization of microbial-cells by semiconductor powders. Fems Microbiology Letters 29:211-214. Maynard, A. D., P. A. Baron, M. Foley, A. A. Shvedova, E. R. Kisin, and V. Castranova. 2004. Exposure to carbon nanotube material: Aerosol release during the handling of unrefined single-walled carbon nanotube material. Journal of Toxicology and Environmental Health-Part A 67 :87-1 07. McWilliams, A. 2006. Nanotechnology: a realistic market assessment. http://www.bccresearch.com/report/NAN03 l B.html. Mitoraj, D., A. Janczyk, M. Strus, H. Kisch, G. Stochel, P. B. Heczko, and W. Macyk. 2007. Visible light inactivation of bacteria and fungi by modified titanium dioxide. Photochemical & Photobiological Sciences 6:642-648. Miyako, E., H. Nagata, K. Hirano, Y. Makita, K. I. Nakayama, and T. Hirotsu. 2007. Near-infrared laser-triggered carbon nanohoms for selective elimination of microbes. Nanotechnology l8:-. Mohanty, B., A. K. Verma, P. Claesson, and H. B. Bohidar. 2007. Physical and anti-microbial characteristics of carbon nanoparticles prepared from lamp soot. Nanotechnology 18:-. Morones, J. R., J. L. Elechiguerra, A. Camacho, K. Holt, J. B. Kouri, J. T. Ramirez, and M. J. Yacaman. 2005. The bactericidal effect of silver nanoparticles. Nanotechnology 16:2346-2353. Mortimer, M., K. Kasemets, M. Heinlaan, I. Kurvet, and A. Kahru. 2008. High throughput kinetic Vibrio fischeri bioluminescence inhibition assay for study of toxic effects of nanoparticles. Toxicology in Vitro 22: 1412-1417. Murr, L. E., E. V. Esquivel, J. J. Bang, G de la Rosa, and J. L. Gardea-Torresdey. 2004. Chemistry and nanoparticulate compositions of a 10,000 year-old ice core melt water. Water Research 38:4282-4296. 28 82. 83. 84. 85. 86. 87. 88. 89. Nadtochenko, V., N. Denisov, O. Sarkisov, D. Gumy, C. Pulgarin, and J. Kiwi. 2006. Laser kinetic spectroscopy of the interfacial charge transfer between membrane cell walls of E-coli and Ti02. Journal of Photochemistry and Photobiology a-Chemistry 181:401-407. Nair, S., A. Sasidharan, V. C. Divya Rani, D. Menon, S. Nair, K. Manzoor, and S. Raina. 2008. Role of size scale of ZnO nanoparticles and microparticles on toxicity toward bacteria and osteoblast cancer cells. Journal of Materials Science. Materials in Medicine. Nel, A., T. Xia, L. Madler, and N. Li. 2006. Toxic potential of materials at the nanolevel. Science 311:622-627. Nepal, D., S. Balasubramanian, A. L. Simonian, and V. A. Davis. 2008. Strong antimicrobial coatings: Single-walled carbon nanotubes armored with biopolymers. Nano Letters 8: 1 896-1901 . Nohynek, G. J., J. Lademann, C. Ribaud, and M. S. Roberts. 2007. Grey goo on the skin? Nanotechnology, cosmetic and sunscreen safety. Critical Reviews in Toxicology 37:251-277. Nyberg, L., R. F. Turco, and L. Nies. 2008. Assessing the impact of nanomaterials on anaerobic microbial communities. Environmental Science & Technology 42: 1938-1943. Oberddrster, G., A. Maynard, K. Donaldson, V. Castranova, J. Fitzpatrick, K. Ausman, J. Carter, B. Karn, W. Kreyling, D. Lai, S. Olin, N. Monteiro-Riviere, D. Warheit, H. Yang, and I. R. F. R. S. I. N. T. S. W. Group. 2005. Principles for characterizing the potential human health effects from exposure to nanomaterials: elements of a screening strategy. Paricle and Fibre Toxicology 2:1-35. Oh, S. D., B. S. Byun, S. Lee, S. H. Choi, M. 1. Kim, and H. G. Park. 2007. Radiolytic synthesis of ag-loaded polystyrene (Ag-PS) nanoparticles and their antimicrobial efficiency against Staphylococcus aureus and Klebsiella pneumoniase. Macromolecular Research 15:285-290. 29 90. 91. 92. 93. 94. 95. 96. 97. 98. Pal, S., Y. K. Tak, and J. M. Song. 2007. Does the antibacterial activity of silver nanoparticles depend on the shape of the nanoparticle? A study of the gram-negative bacterium Escherichia coli. Applied and Environmental Microbiology 73: 1712-1720. Rahman, Q., J. Norwood, and G Hatch. 1997. Evidence that exposure of particulate air pollutants to human and rat alveolar macrophages leads to differential oxidative response. Biochemical and Biophysical Research Communications 240:669-672. Reddy, K. M., K. Feris, J. Bell, D. G Wingett, C. Hanley, and A. Punnoose. 2007. Selective toxicity of zinc oxide nanoparticles to prokaryotic and eukaryotic systems. Applied Physics Letters 90:-. Reijnders, L. 2006. Cleaner nanotechnology and hazard reduction of manufactured nanoparticles. Journal of Cleaner Production 14: 124-133. Rincon, A. G, and C. Pulgarin. 2004. Bactericidal action of illuminated Ti02 on pure Escherichia coli and natural bacterial consortia: post-irradiation events in the dark and assessment of the effective disinfection time. Applied Catalysis B-Environmental 49:99-1 l2. Rincon, A. G, and C. Pulgarin. 2004. Effect of pH, inorganic ions, organic matter and H202 on E—coli K12 photocatalytic inactivation by Ti02 - Implications in solar water disinfection. Applied Catalysis B-Environmental 51:283-302. Saito, T., T. Iwase, J. Horie, and T. Morioka. 1992. Mode of photocatalytic bactericidal action of powdered semiconductor Ti02 on mutans Streptococci. Journal of Photochemistry and Photobiology B-Biology 14:369-379. Sawai, J., H. Igarashi, A. Hashimoto, T. Kokugan, and M. Shimizu. 1995. Effect of ceramic powder slurry on spores of Bacillus subtilis. Journal of Chemical Engineering of Japan 28:556-561. Sawai, J., H. Igarashi, A. Hashimoto, T. Kokugan, and M. Shimizu. 1996. Effect of particle size and heating temperature of ceramic powders on antibacterial activity of their slurries. Journal of Chemical Engineering of Japan 30 99. 100. 101. 102. 103. 104. 105. 106. 29:251-256. Sawai, J., E. Kawada, F. Kanou, H. Igarashi, A. Hashimoto, T. Kokugan, and M. Shimizu. 1996. Detection of active oxygen generated from ceramic powders having antibacterial activity. Journal of Chemical Engineering of Japan 29:627-633. Sayes, C. M., J. D. Fortner, W. Guo, D. Lyon, A. M. Boyd, K. D. Ausman, Y. J. Tao, B. Sitharaman, L. J. Wilson, J. B. Hughes, J. L. West, and V. L. Calvin. 2004. The differential cytotoxicity of water-soluble fiillerenes. Nano Letters 4:1881-1887. Sayes, C. M., R. Wahi, P. A. Kurian, Y. P. Liu, J. L. West, K. D. Ausman, D. B. Warheit, and V. L. Colvin. 2006. Correlating nanoscale titania structure with toxicity: A cytotoxicity and inflammatory response study with human dermal fibroblasts and human lung epithelial cells. Toxicological Sciences 92:174-185. Sondi, 1., and B. Salopek-Sondi. 2004. Silver nanoparticles as antimicrobial agent: a case study on E-coli as a model for Gram-negative bacteria. Journal of Colloid and Interface Science 275: 177-182. Stoimenov, P. K., R. L. Klinger, G L. Marchin, and K. J. Klabunde. 2002. Metal oxide nanoparticles as bactericidal agents. Langmuir 18:6679-6686. Sunada, K., T. Watanabe, and K. Hashimoto. 2003. Studies on photokilling of bacteria on Ti02 thin film. Journal of Photochemistry and Photobiology a-Chemistry 156:227-233. ' Thill, A., O. Zeyons, O. Spalla, F. Chauvat, J. Rose, M. Auffan, and A. M. Flank. 2006. Cytotoxicity of CeOz nanoparticles for Escherichia coli. Physico-chemical insight of the cytotoxicity mechanism. Environmental Science & Technology 40:6151-6156. Tong, Z. H., M. Bischoff, L. Nies, B. Applegate, and R. F. Turco. 2007. Impact of fullerene (C-60) on a soil microbial community. Environmental Science & Technology 41 :2985-2991. 31 107. 108. 109. 110. 111. 112. 113. 114. 115. Tsao, N., P. P. Kanakamma, T. Y. Luh, C. K. Chou, and H. Y. Lei. 1999. Inhibition of Escherichia coli-induced meningitis by carboxyfullerence. Antimicrobial Agents and Chemotherapy 43:2273-2277. Tsao, N., T. Y. Luh, C. K. Chou, T. Y. Chang, J. J. Wu, C. C. Liu, and H. Y. Lei. 2002. In vitro action of carboxyfullerene. Journal of Antimicrobial Chemotherapy 49:641 -649. Tsao, N., T. Y. Lub, C. K. Chou, J. J. Wu, Y. S. Lin, and K. Y. Lei. 2001. Inhibition of group A streptococcus infection by carboxyfullerene. Antimicrobial Agents and Chemotherapy 45:1788-1793. Turk, V., C. Kaiser, and S. Scballer. 2008. Invisible but tangible? Societal opportunities and risks of nanotechnologies. Journal of Cleaner Production 16:1006-1009. Wakefield, G, M. Green, S. Lipscomb, and B. Flutter. 2004. Modified titania nanomaterials for sunscreen applications - reducing free radical generation and DNA damage. Materials Science and Technology 20:985-988. Wamer, W. G, J. J. Yin, and R. R. Wei. 1997. Oxidative damage to nucleic acids photosensitized by titanium dioxide. Free Radical Biology and Medicine 23:851-858. Wang, J. J., B. J. S. Sanderson, and H. Wang. 2007. Cyto- and genotoxicity of ultrafine Ti02 particles in cultured human lymphoblastoid cells. Mutation Research-Genetic Toxicology and Environmental Mutagenesis 628:99-106. Warheit, D. B., T. R. Webb, K. L. Reed, S. Frerichs, and C. M. Sayes. 2007. Pulmonary toxicity study in rats with three forms of ultrafine-TiO2 particles: Differential responses related to surface properties. Toxicology 230:90-104. Warheit, D. B., T. R. Webb, C. M. Sayes, V. L. Colvin, and K. L. Reed. 2006. Pulmonary instillation studies with nanoscale Ti02 rods and dots in rats: Toxicity is not dependent upon particle size and surface area. Toxicological Sciences 91:227-236. 32 116. 117. 118. 119. 120. 121. 122. 123. 124. Wei, C., W. Y. Lin, Z. Zainal, N. E. Williams, K. Zhu, A. P. Kruzic, R. L. Smith, and K. Rajeshwar. 1994. Bactericidal activity of Ti02 photocatalyst in aqueous-media - toward a solar-assisted water disinfection system. Environmental Science & Technology 28:934-938. Williams, D. N., S. H. Ehrman, and T. R. P. Holoman. 2006. Evaluation of the microbial growth response to inorganic nanoparticles. Journal of Nanobiotechnology 4. Wolfrum, E. J., J. Huang, D. M. Blake, P. C. Maness, Z. Huang, J. Fiest, and W. A. Jacoby. 2002. Photocatalytic oxidation of bacteria, bacterial and fungal spores, and model biofilm components to carbon dioxide on titanium dioxide-coated surfaces. Environmental Science & Technology 36:3412-3419. Xu, X. H. N., W. J. Brownlow, S. V. Kyriacou, Q. Wan, and J. J. Viola. 2004. Real-time probing of membrane transport in living microbial cells using single nanoparticle optics and living cell imaging. Biochemistry 43:10400-10413. Yamamoto, O. 2001. Influence of particle size on the antibacterial activity of zinc oxide. International Journal of Inorganic Materials 3:643-646. Yang, Z. H., and C. S. Xie. 2006. Zn2+ release from zinc and zinc oxide particles in simulated uterine solution. Colloids and Surfaces B-Biointerfaces 47: 140-145. Zhang, L. L., Y. H. Jiang, Y. L. Ding, M. Povey, and D. York. 2007. Investigation into the antibacterial behaviour of suspensions of ZnO nanoparticles (ZnO nanofluids). Journal of Nanoparticle Research 9:479-489. Zhang, L. L, J. C. Yu, H. Y. Yip, Q. Li, K. W. Kwong, A. W. Xu, and P. K. Wong. 2003. Ambient light reduction strategy to synthesize silver nanoparticles and silver-coated Ti02 with enhanced photocatalytic and bactericidal activities. Langmuir 19:10372-10380. Zheng, H., P. C. Maness, D. M. Blake, E. J. Wolfrum, S. L. Smolinski, and W. A. Jacoby. 2000. Bactericidal mode of titanium dioxide photocatalysis. Journal of Photochemistry and Photobiology a-Chemistry 130:163-170. 33 CHAPTER 2 TOXICITY OF OXIDE NANOPARTICLES ON ESCHERICHIA C 0L1 AND EXI G U OBA C TERI UM SIBIRIC UM 34 Abstract NPs are being used in a variety of areas, including cosmetics, pigments, water purification, environmental remediation and drug delivery. However, their toxicological and environmental effects have not been. thoroughly investigated. In our studies, zinc oxide (ZnO) and titanium oxide (Ti02) NPs were found to decrease cell viability of Escherichia coli strain K12 (Gram-negative) and Exiguobacterium sibiricum strain 255-15 (Gram-positive) in a dose-dependent manner, while silicon oxide (Si02) and iron oxide (Fe3O4) NPs did not inhibit the microbial growth after a 12 h exposure. Five nanometer Ti02 NPs were more toxic to both E. coli and E. sibiricum than 150 nm Ti02 NPs, which confirmed that size is one of the most important factors for NP toxicity. ZnO NPs displayed more toxicity to E. coli compared to E. sibiricum, while Ti02 NPs were more toxic to E. sibiricum than to E. coli. Zn2+ ion dissolved from ZnO NPs in both LB and TSB medium, and reached saturation of approximately 100- 110 ppm. E. coli was more vulnerable to the toxicity of Zn2+ than E. sibiricum. Transmission electron microscopy (TEM) and energy dispersive spectroscopy (EDS), showed that Si02 and Ti02 NPs were internalized by E. sibiricum but not by E. coli, and ZnO NPs were found in both. We concluded that size and chemical composition play a key role in NP toxicity, and that solubility of metal ions from NPs could also attribute to the toxicity of NPs. These results also show that different bacteria have different tolerance and internalization ofNPs. 35 Introduction NMs are defined by the US. National Nanotechnology Initiative as materials that have at least one dimension in the 1 to 100 nm range (34). Nanotechnology has become one of the rapidly developing fields with the potential to revolutionize medical science, biology and engineering. Meanwhile, the concerns about the toxic and environmental impacts of exposure to these NMs have been increasing and starting to attract public and governmental attention. The critical risk assessment issues about NMs include toxicology, exposure assessment, environmental and biological fate, persistence, sustainability, transportation and transformation. Moreover, a number of studies suggest that NMs might cause damage, such as lung cancer, silicosis and some autoimmune diseases (9, 10, 13,20, 37, 42). NPs, the type most important in NMs, are applied in environmental remediation, catalysis, drug delivery, cosmetics, water purification and lubrication (2, 14, 18, 33, 49). $02, Ti02 and ZnO NPs were the main NPs test in this study. Si02 is used as an additive in food, drugs, cosmetics and printer toners, as well as it is a good semiconductor and an electrical insulator. Ti02 is used‘as a photocatalyst for disinfection of bacteria and a number of the papers explored the mechanism (3, 7, 27, 29, 38, 40, 44). It is also employed as a pigment in paints and as a sunscreen product for cosmetics. ZnO is widely used for application as pigments, semiconductors, and as a sunscreen. It was estimated that globally 1000 tons of NPs were used for sunscreen products during 2003 and 2004 and more than 2500 tons in 2008 (4). Because of their numerous applications, NPs may 36 enter the environment during their life cycle and end up as environmental contaminants in water, air or soil. However, very few studies on NPs about their environmental and toxicological effects on the microorganisms have been conducted, and their mechanism of toxicity is still unknown. The main properties of NPs that may affect the toxicity include size distribution, agglomeration state, shape, crystal structure, chemical composition, surface area, surface chemistry, surface charge and porosity (34). We evaluated the effects of Si02, Fe3O4, Ti02 and ZnO NPs on E. coli K12 and E. sibiricum representative Gram—negative and Gram-positive bacteria, and explored if the size and composition of these NPs influenced the degree of toxicity. Moreover, we investigated the dissolution of ZnO NPs, and compared the toxicity of Zn2+ ion with that of the NP suspension. TEM and EDS were employed to study the internalization of NPs and the morphological changes of these bacteria after the exposure to NPs. Materials and Methods Reagents and bacterial culture. E. coli K12 was grown aerobically in Luria-Bertani (LB) liquid broth at 37 °C and 200 rpm for 12 h. E. sibiricum was cultured in tryptic soy broth (TSB) medium at 30 °C and 150 rpm for 12 h. Solid medium for E. coli or E. sibiricum was as above with 1.5% agar (Difco). The bacteria were in stationary phase when the NPs were added, as well as after 12 h exposure. All the cultures with NPs were grown under the same conditions. Preparation of NP suspensions. Si02 (10 nm and 1 pm), Fe3O4 (50 nm and 1 um), 37 Ti02 (5 nm, 150 nm and 5 mm) and ZnO (60 nm and 1 mm) were purchased from Sigma-Aldrich (St. Louis, MO). Sixty namometer Si02 was synthesized by reverse microemulsion (45, 50). NPs stock solutions of 100 g/L were prepared in deionized and distilled water. Assessment of NP toxicity to bacteria. E. coli K12 and E. sibiricum were cultured as above for 12 h with NPs at 0, 10, 50, 100, 500, 1000, 2000, and 5000 ppm. Cultures were then diluted and spread onto LB or trypic soy agar (TSA) plates, and colonies were counted after growth over-night. All the cultures were grown in triplicate. Evaluation of Zn2+ toxicity in ZnO NP medium. Different concentrations of 60 nm and 1 pm ZnO NPs were mixed in TSB and LB for 12 h (50 ppm, 200 ppm, 1000 ppm and 5000 ppm). After 20 min of centrifugation at 20,000Xg, the supernatants were collected, diluted and measured with atomic absorption spectrometry. Zinc standard solution was purchased from Fischer Scientific Co. (Pittsburgh, PA) and diluted to construct a standard curve for Zn2+ concentration. E. coli K12 and E. sibiricum were cultured in triplicate in each of three concentrations of ZnClz (0.5 mM, 1.0 mM and 2.0 mM). Culture viability was measured by plate counts and all the samples were prepared in triplicate. Toxicity of NPs versus soluble metals from NPs. LB or TSB medium containing possible metal impurities from TiOz or ZnO NPs was prepared by making suspension of NPs with concentrations of 1000, 2000, and 5000 ppm for 10 nm, 150 nm and 5 pm Ti02 NPs, and, 100, 200, 500, 1000, 2000, and 5000 ppm for 60 nm ZnO NPs. After leaving 38 the LB suspension at 30 °C and TSB suspension at 37 °C for 12 h, the NPs were removed by centrifuging the suspension at 20,000Xg for 20 min. The supernatant was collected and used to culture E. coli and E. sibiricum respectively for 12 h, and the microbial growth was evaluated by plate counting method. Morphological studies E. coli K12 and E. sibiricum with 1000 ppm of NPs (60 nm Si02, 5 nm Ti02 and 60 nm ZnO) were cultured in LB or TSB medium overnight at 30 0C or 37 0C respectively. After culturing, the mixture of cells and NPs in each sample was collected by centrifugation at 12,000 X g for 30 min. These samples were fixed with 2.5% glutaraldehyde and 2.5% paraformaldehyde in 0.1 M cacodylate buffer and then 1% osmium tetroxide in cacodylate buffer, dehydrated in serial acetone buffer (from 30% to 100%), and infiltrated with acetone: Poly/Bed 812 buffer for TEM. The samples were cut by MTX ultramicrotome into 70-100 nm thick sections. All the sections were collected on copper grids, and a JEOL 2200F S Transmission Electron Microscope was used for the EDS-TEM detection. Results Antibacterial activity of ZnO and Ti02 NPs. E. coli and E. sibiricum was exposed to two different sizes of ZnO NPs for 12 h to evaluate their effect on microbial growth. As illustrated in Figure 2.1, the toxicity of ZnO NPs to E. coli and E. sibiricum rose with the increasing doses. ZnO NPs significantly inhibited the growth of E. coli when the 39 concentration was higher than 100 ppm (Figure 2.1(a)). The 60 nm ZnO NPs were more toxic to E. coli than the 1 pm ones. In the case of E. sibiricum, a gradual increase in growth inhibition was observed with increasing doses of ZnO NPs, while E. sibiricum showed more tolerance to ZnO NPs (Figure 2.1(b)). In addition, the 60 nm ZnO NPs displayed less toxicity to E. sibiricum than the 1 pm NPs, which was different from the size-effect of ZnO NPs on E. coli. Similar to the results with ZnO NPs, all three Ti02 NPs showed a gradual increase in toxicity to both E. coli and E. sibiricum as the doses increased (Figure 2.2). Moreover, E. sibiricum was more vulnerable to Ti02 NPs than E. coli, which was contrary to the toxicity of ZnO NPs to E. coli and E. sibiricum. In the case of E. coli, both 5 nm and 150 nm NPs significantly inhibited growth of E. coli at 2000 and. 5000 ppm, with 5 nm NPs being more toxic than the 150 nm ones (Figure 2.2(a)). The 5 pm Ti02 NPs did not display any toxicity even at 5000 ppm. As illustrated in Figure 2.2(b), the growth of E. sibiricum was inhibited significantly by all the three different size Ti02 NPs, even for 5 pm NPs when their concentrations were higher than 1000 ppm. The 5 nm NPs strongly inhibited the growth of E. sibiricum at concentrations above 500 ppm while the 150 nm TiOz NPs were intermediate in their inhibition of the growth. Zn2+ toxicity in ZnO NP solution. Zn2+ ion concentration in LB or TSB solution of ZnO NPs was measured with atomic absorption spectrometry, and the results demonstrated that the concentrations of Zn2+ ion rose with the increasing ZnO NP concentrations (Table 2.1). The concentrations of Zn2+ ion dissolved from the same 40 concentration of 60 nm and 1 pm ZnO NPs were similar, which suggested that the size of NPs did not influence the dissolution of ZnO NPs much. Additionally, there was no significant difference in Zn2+ ion concentrations between the same concentration of ZnO NPs in LB and TSB broth. In the ZnO NP suspension of 50 ppm, the concentration of Zn2+ ion was approximately 40 ppm in either TSB or LB broth, which meant that most of ZnO NPs were dissolved in the solution. The concentration of Zn2+ was saturated at 100- 110 ppm in TSB or LB medium, which occurred when the concentration of NPs was higher than 200 ppm. In Table 2.2, the viability 0f E. coli or E. sibiricum decreased with increasing concentrations of ZnClz in broth, with E. coli being more vulnerable of the two strains. The growth of these two strains to Zn2+ ions suggested that the most of overall ZnO NP toxicity came from the dissolved Zn2+. The ZnO NPs, however, still played an important role in the toxicity for the solutions containing high concentrations of ZnO NPs, since the concentration of Zn2+ reached the saturation of 110 ppm. Toxicity of NPs versus soluble metals from NPs. The toxicity of Ti02 or ZnO NPs was compared to that of possibly released metal ion Ti4+ or Zn2+ from the same concentration of suspension. Since no toxicity in the Ti02 NP suspension lower than 1000 ppm was found for either E. coli or E. sibiricum, only the suspensions > 1000 ppm were tested. The soluble fraction of 10 nm, 150 nm and 5 pm Ti02 NPs showed little effect on the grth of E. coli (Figure 2.5(a)). The reason for this is probably that the Ti02 NPs is highly insoluble, so the concentration of Ti“ ion is not high enough to cause damage to E. 41 coli. Another reason may be that membrane of E. coli is not that permeable to Ti02 NPs compared to E. sibiricum. In the case of E. sibiricum (Figure 2.5(b)), the biggest difference between the suspension and the soluble metal was seen at 5 um Ti02 NPs: no toxicity was found in soluble fractions while the viability of E. sibiricum in suspension of 2000 ppm and 5000 ppm 5 pm Ti02 NPs was accordingly 74 % and 20 %. Compared to Ti02 NPs, the soluble fraction in ZnO NP suspension played a more important role in the microbial growth. For E. coli, the suspension of ZnO NPs was 10 to 40 fold more toxic than the soluble fraction, when the ZnO NP concentrations ranged from 100 to 200 ppm (Figure 2.6(a)). Furthermore, there was no toxicological difference on the microbial grth between the suspension and the soluble fraction at the concentration of ZnO NPs higher than 500 ppm. Compared to E. coli, E. sibiricum was less vulnerable to Zn2+ ion or ZnO NPs. Contrary to the results with E. coli, there was almost no significant toxic difference on the growth of E. sibiricm between the ZnO NP suspension and the soluble fraction, when the ZnO NP concentration is no higher than 1000 ppm (Figure 2.6(b)). At high concentrations of ZnO NPs (2000 and 5000 ppm), however, the soluble fraction was 3 to 6 fold less toxic than the whole suspension. It is well known that low concentration of Zn2+ is safe or even beneficial for the bacteria, by providing stability to the membrane for example. For E. coli, the concentration of Zn2+ was unsaturated and not that high in solutions containing low concentrations of ZnO NPs, so ZnO NPs in the suspension could raise the toxicity and played a predominant role on the inhibition of the microbial growth. Furthermore, when 42 the concentration of Zn2+ reached saturation at the high concentrations of ZnO NP suspension, the soluble fraction dominated the toxicological effects. Compared to E. coli, E. sibiricum was more tolerant to the toxicity of Zn2+. That is why the increasing concentrations of ZnO NPs in the suspension still could enhance the inhibition to the growth of E. sibiricum when the concentration of Zn2+ was saturated. Morphological studies. TEM was applied to visualize the morphological changes in E. coli and E. sibiricum exposed to 1000 ppm NPs in LB or TSB medium for 12 h. EDS was performed for the validation of NPs, because osmium tetroxide, used to fix lipid in the membrane, gave some confusing black spots inside cells. The 60 nm 810; NPs were found in E. sibiricum not in E. coli, which meant that it was easier for Si02 NPs to go through the membrane of E. sibiricum than that of E. coli (Figure 2.7(a)). The EDS results revealed that Silicon element appeared along the cross-section line and there was an increase in the distribution of both silicon and oxygen elements along the line, which confirmed the internalization of Si02 NPs (Figure 2.7(b) and (c)). Furthermore, the width of the silicon peak was similar to the 60 nm size of NPs. Similar to the results with ZnO NPs, 5 nm Ti02 NPs was found inside E. sibiricum cells but none in E. coli (Figure 2.8(a)). This is consistent with the toxicological results that Ti02 NPs showed more toxic to E. sibiricum than to E. coli. The EDS results confirmed the presence of elemental titanium along the cross-section line that corresponds on the thin section samples. The width of the titanium peak was similar to the size of the NPs in the cell (Figure 2.8(b)). 43 Sixty nanometer ZnO NPs were found inside both E. coli and E.sibiricum cells (Figure 2.9(a) and (c)), which suggested that the ZnO could pass the both of membranes. Lots of dead cells with broken membranes and leaking cytoplasm were also found under TEM. The EDS data confirmed that the region along the cross-section line contained elemental zinc. The distribution of zinc displayed a peak whose width was similar in size to ZnO NPs (Figure 2.9(b) and ((1) dark spot). Together these data validated that the dark spot is a ZnO NP. Similar internalization results was published by Brayner et al. (5). Discussion The chemical composition of NPs is one of the most important factors determining biological response, and causes their toxicity. As illustrated in Figure 2.1 and 2.2, ZnO NPs and Ti02 NPs were toxic to both E. coli and E. sibiricum, while Si02 and Fe3O4 NPs were not generally. Only in the case of E. sibiricum, did 500 ppm 10 nm Si02 NPs inhibit growth, and then only after the exposure time was increased to 48 h. The composition of NPs also influences the possible mechanisms of NP toxicity. For instance, ZnO NPs are slightly soluble (5.4mg/L), and the released Zn2+ ion plays a key role in ZnO NP toxicity. In the case of Ti02 NPs, the ROS leads to oxidation of membrane and proteins, and hinders nutrient uptake and respiration, resulting cell death. Recently several papers show the toxicity of Si02 NPs to E. coli, B. subtilis and human cells (1, 21 , 24), and some groups reported that no cytotoxicity was found for iron oxide NPs, similar to our results (15, 16, 19, 23, 36). All these data demonstrate that chemical 44 composition is one of the physicochemical properties important to NP toxicity. Size is also an important factor to NP toxicity. As size decreases, both surface area and surface reactive groups increase, which may enhance the NP toxicity (32, 35). We found that the smaller ZnO or Ti02 NPs were more toxic to E. coli than larger NPs (Figure 21(3) and 2.2(a)), similar to previous findings (19, 51). There were some conflicts in the size-effect with E. sibiricum, since 1 um ZnO NPs were more toxic than 60 nm ZnO NPs (Figure 2.1(b)), and 5 um Ti02 NPs were more toxic to E. sibiricum than 150 nm Ti02 NPs (Figure 2.2(b)). Aggregation of NPs in the medium may be one reason for these conflicts. High concentrations of salts and proteins in the medium can cause the aggregation of NPs, which increases the size the cell experienced (26). For instance, Long et al. found that commercial grade, nano-size Ti02 NPs aggregated in Hank’s basic salt solution and Dulbecco’s modified eagle’s medium, and the hydrodynamic diameter of the TiOz NPs increased to 2368i163 nm at 120 ppm in their studies (25). Although there is no clear consensus about the upper size limit for the NP toxicity, it was estimated that it might be between 65 nm and 200 nm (11). Additionally, there are other factors of NPs which can influence NP toxicity, such as shape and surface modification (41). So it is important for toxicology and exposure assessment of NPs to determine and characterize these parameters. Gram-positive or Gram-negative species may show different tolerance to the toxicity of the same NPs. For example, Ti02 NPs were more toxic to E. sibiricum than to E. coli 45 (Figure 2.2). Some groups reported that the Ti02 NPs should be more toxic to Gram negative than Gram positive species, since Gram positive species have thicker cell walls and can form spores (49). E. sibiricum is non-spore-forming, which may be one of the reasons why Ti02 NPs are more toxic to E. sibiricum. We also found that ZnO NPs were more toxic to E. coli compared to E. sibiricum. Adams et al. (1) found that ZnO NPs were the most toxic ones of three NMs they tested, and were more toxic to B. subtilis, Gram positive, than to E. coli, Gram negative. Other external factors, such as the initial concentration of bacteria, the physiological state of the bacteria, light intensity, pH and 02, also can influence the toxicity of NPs (38). The dissolution of NPs was hypothesized as one of the mechanisms by which NPs interact with cells (32). The dissolution can be influenced by the coating of NPs, pH of solution and other factors (12, 52). Ti02 is highly insoluble (0.15 mg/L) while ZnO is slightly soluble (5.4 mg/L) (6). Even if Ti4+ is dissolved into solution, the reaction below usually happens: Ti4+ + 2HZO —+ Tio2 + 4H+ Ti2+ is much less stable than Ti“, and it is probably oxidized by 02 to form Ti“. The dissolution of ZnO NPs was previously reported, with the suggestion that toxicity might depend on Zn2+ ion in NP solution (6, 12, 17, 22). The dissolution of ZnO NPs may worsen the adverse effects on aquatic biota and cause more ecotoxicity (39). We did not find that the dissolution of ZnO NPs was size-dependent, although it had been reported that Zn2+ release rate of ZnO NPs was faster than that of ZnO micro-particles (52). Zinc 46 ions play an important role in cellular process, mostly as an enzyme cofactor (47). It has been reported that high concentrations of Zn2+ ion are harmful for cells by inhibiting some enzyme activity or affecting ATP (8, 28, 48). The viability results demonstrated that Zn2+ was more toxic to E. coli than E. sibiricum. According to the genome sequences of E. coli and E. sibiricum in the database of National Center for Biotechnology Information, there are more proteins annotated for zinc transport, efflux, or uptake in E. coli than in E. sibiricum, which may make E. coli more vulnerable to Zn2+ toxicity. Intemalization of NPs can be influenced by size and chemical composition of NPs and by types of microorganisms. For instance, 5 nm Ti02 NPs were found in E. sibiricum rather than in E. coli, which suggests that it is easier for Ti02 NPs to go through the membrane of E. sibiricum than that of E. coli. This could be one of the reasons why Ti02 NPs were more toxic to E. sibiricum than E. coli, since the internalized NPs may cause direct damage and induce serious toxicity. Ti02 NPs can decompose and peroxidate membranes, which is the explanation for the killing effect by Ti02 photocatalysis (27, 40, 44). In the ZnO NP samples, many dead cells with damaged membranes and leaking intercellular contents were noted, also observed in a previous study (5). Furthermore, we investigated the morphology of E. coli and E.sibiricum exposed to 50 nm Fe3O4 NPs and 5 pm Ti02 particles by TEM, and found that both E. coli and E. sibiricum could internalize Fe3O4 NPs, although Fe304 NPs were non-toxic. No 5 pm Ti02 particles were found inside these microorganisms. The largest Ti02 particles found inside the cells were around 150 nm. Other NPs, such as CeOz and silver, were previously found internalized 47 in E. coli or incorporated into the membrane, as well as interfacial charge transfer and the interaction between the Ti02 NPs and membrane of E. coli (30, 43, 46). There is no clear explanation, however, about the mechanisms of internalization. The contact between NPs and bacteria was enhanced by the repulsive force generated between negatively charged bacterial cells and positively charged NPs (31). In conclusion, we demonstrated that ZnO and Ti02 NPs were toxic to E. coli and E. sibiricum in a dose-dependent manner, while SiOz and Fe3O4 NPs did not inhibit their growth during a 12 h exposure. Among these NPs, ZnO NPs were the most toxic ones for either of these two microorganisms. ZnO was more toxic to E. coli than to E. sibiricum with the toxicity mostly due to the Zn2+ ion dissolved from the NP, while TiOz was more toxic to E. sibiricum than to E. coli. By TEM-EDS, we found Si02 and Ti02 NPs internalized only in E. sibiricum not in E. coli, while ZnO NPs appeared inside both of these bacteria. 48 1000 - - 60 nm 100 1 um " ii u: g C E e\: 10 J I *Itbl 11* go 5 l u 3 1 , "' _ a 8 i an: a E 11 u 1* ** 13* *1! L ‘ a It ~— slut: ‘ - ** *t ** 01 l i i .. 1". :5 a 'i if": E 5: is 0.01 " :91 l‘; i“ 3 g iii A 3: control 50 100 200 300 400 500 1000 2000 5000 Nanoparticles concentration (ppm) #9533- b 1000 - 1 um - 60 nm 100 " r: " at it: to, B a an at g; 11 1:33 i it 1 V ,, .' .' . E u: an 3:. .- :2 ii . ; 9 1o - l. g iii a: .. E g 31': I' l :51} ”'- link 8 ii a s ' . Z 1 F 1‘ E a 1 - it -. 1:. ' l . “ g n 3"; ‘51 i t. E t 5’fo / 0.1 - a control 50 100 200 300 400 500 1000 2000 5000 Nanoparticle concentration (ppm) Figure 2.1 (a) The effect of ZnO NPs on E. coli. (b) The effect of ZnO NPs on E. sibiricumi. *p<0.05 versus control cells;*"'p<0.005 versus control. 49 140 - 120 - 100 a 80+ 60- Cell Growth (%) 40‘ 20a 120 a 100 - 80- 60- Cell Growth (%) 4o- 20' 1." ' -w.lmmme ' :- control 3 ,l WHJWW control 32' 5O j_l if t! 50 -5nm - 150 nm - 5 um i a ! ‘.‘ * l i *- -' p . an: : ‘ fl } 1 f a l l " i ii I .~ i: “ ‘ l ~ 1 t 100 500 1 000 2000 5000 TiO2 concentration (ppm) b — 5 nm T - 150 nm 1‘ - 5 um i at: l * . if 1 . t l a , ii , t. l ,i ,‘ ‘7' l i ll . . l; f .. - 100 500 1000 2000 5000 Ti02 concentration (ppm) Figure 2.2 (a) The effect of Ti02 NPs on E. coli. (b) The effect of Ti02 NPs on E. sibiricum. *p<0.05 versus control cells;**p<0.005 versus control. 50 - 60nm 140 m1um Cell Growth (%) is 20‘ control 50 100 Fe304 concentration (ppm) 140 - - 60 nm 1 l l 60- Cell growth (%) 4o — - it . 5 in .+ . 20‘ Drag: ‘ control 50 100 500 1 000 2000 5000 Fe304 concentration (ppm) Figure 2.3 (a) the toxicity of 60 nm and 1 pm iron oxide NPs to E coli, (b) the toxicity of 60 nm and 1 pm iron oxide NPs on E. sibiricum. 51 140 ‘ - 10 nm _ 1 um 120 - 100 ‘ .‘133 {a {3 o z .c 80 ‘ g E 9 ‘_ 2’ 60 - ; 5 8 5' 2:1 40 ‘ ~ 3: 20 - 1; 0113'"; § Fifi 2‘5». control 50ppm 100ppm 500ppm 1000ppm 2000ppm 5000ppm nanoparticles concentration (ppm) b 140 - - NPs1O nm Es: NPs1 um 120 - 100 J E El .5 E 80 d " -.; 2 f"- I 52 6O 4 '. -.-' L. t 35 I . 40 ‘ :' fir % l. .‘ ' 20 ‘ a control 50ppm 100ppm 500ppm 1000ppm 2000ppm 5000ppm Nanoparticle Concentration (ppm) Figure 2.4 (a) the influence of Si02 NPs (10 nm and 1 pm) on the growth of E. coli, (b) the influence of Si02 NPs (10 nm and 1 pm) on the growth of E. sibiricum. 52 Table 2.1 The concentration of Zn2+ ion dissolved in LB or TSB from various concentrations of ZnO NPs ZnO NPs Zn2+ Conc in LB (ppm) Zn2+ Come in TSB (ppm) Cone 60 nm 1pm 60 nm 1pm (ppm) 0 1.2:t0.0 1.2:t0.0 0.7:i:0.0 0.7:t0.0 50 38.0:t2.8 37.0:t:1.4 40.0i0.0 38.5:h2.1 200 90.0:t0.0 85.0:t7.1 110.0i0.0 100.0:t0.0 1000 100.0:t0.0 100.0i0.0 10502157.] 100.0i0.0 5000 115.0i7.1 100.0101) 110.0i0.0 100.0100 Table 2.2 Viability of E. coli and E. sibiricum in various concentrations of Zan solution ZnClz Conc Viability (%) (ppm) E. coli E. sibiricum Control 100.0i10.3 100.0i2.0 32.5 87.3i10.4 113.9i4.6 65 68.2:t4.2 77.3i7.8 130 <0.01 16.4i3.9 53 a + 5 nm Ti02 solution only —0— 5 nm Ti02 NPs and solution 140 1 + 150 nm Ti02 solution only —A— 150 nm Ti02 NPs and solution + 5 um Ti02 solution only 120 ' —U—- 5 um Ti02 NPs and solution g 100 - g '5 to '5 80 - 60 d 40 l I I l control 1 000 2000 5000 TiO2 NP concentration (ppm) + 5 nm Ti02 solution only —0— 5 nm Ti02 NPs and solution + 150 nm Ti02 solution only —A— 150 nm Ti02 NPs and solution 140 ' + 5 um Ti02 solution only —t:l— 5 urn Ti02 NPs and solution 120 a 100 - . «b» \. :3? A“ :5 so - .. .0 .9 > 40 a 20 “ 6 I 0 .. : control 1000 ' 2000 5000 TiO2 NPs concentration (ppm) Figure 2.5 (a) The toxicity of Ti02 NP suspension and soluble fraction on E. coli. (b) The toxicity of Ti02 NP suspension and soluble fraction on E. sibiricum. 54 1000 - 100 g + 60 nm ZnO solution only —0- 60 nm ZnO NPs and solution g 10 - g B .9 > 1 r 0.1 d 0.01 r . r r r r f control 100 200 500 1000 2000 5000 ZnO NP concentration (ppm) b 1000 - + 60 nm ZnO solution only —0— 60 nm ZNO NPs and solution 100 4 E g 10 . .o .3! > 1 .. 0.1 I I I T l l I control 100 200 500 1000 2000 5000 ZnO NP concentration (ppm) Figure 2.6 (a) The toxicity of ZnO NP suspension and soluble fraction on E. coli. (b) The toxicity of ZnO NP suspension and soluble fraction on E. sibiricum. 55 dowmxo co 8:353 E .825 5852“. o 5:353 § .._§o.:.s.a.m 5 Now a mo owes Ema 3 sesame 5 ~05 co 5:335 .2 as»; .E 5905 B. S . 80:5 . 8w . SEEN _ E: com 56 .C. go 8855 05 wean—55v xaoq of. .moEEMm co 2.: 8:03-320 2: mac? 5253: .5553 o 8:355 3v £8 55:3: .m 02% oz Neg So owns: Ewe 3 .3352 58:5: .m a No: co 8:355 .3 2%: sage omw Pmawchcozuflw xiii”? . 57 .5520 8a co EHBEWE 9; SEE: .m a 05 2o owaé 2m; 3 e8 a co oaeom 2: 53a 2.: 2: mg: 32% 8o o 8:335 3V :8 .m 5 oz 05 a co owaé SE E .25 83an 28:3,: .m an 3-5 =8 .m a OS co 8:38am .3 2am; 25 5 25 RE Er D rl[1’rrV‘| D .— R 8 o __ . 33E. cozomm .._ Ecoom _ llllvillrvl|lIvvllll'lleIIlvlllll O V 58 References 1. Adams, L. K., D. Y. Lyon, A. McIntosh, and P. J. J. Alvarez. 2006. Comparative toxicity of nano-scale Ti02, Si02 and ZnO water suspensions. Water Science and Technology 54:327-334. 2. Aitken, R. J., M. Q. Chaudhry, A. B. A. Boxall, and M. Hull. 2006. Manufacture and use of nanomaterials: current status in the UK and global trends. Occupational Medicine-Oxford 56:300-306. 3. Block, S. S., V. P. Seng, and D. W. Goswami. 1997. Chemically enhanced sunlight for killing bacteria. Journal of Solar Energy Engineering-Transactions of the Asme 119:85-91. '4. Borm, P. J. A., D. Robbins, S. Haubold, T. Kuhlbusch, H. Fissan, K. Donaldson, R. Schins, V. Stone, W. Kreyling, J. Lademann, J. Krutmann, D. Warheit, and E. Oberdorster. 2006. The potential risks of nanomaterials: a review carried out for ECETOC. Paricle and Fibre Toxicology 3. 5. Brayner, R., R. Ferrari-Iliou, N. Brivois, S. Djediat, M. F. Benedetti, and F. F ievet. 2006. Toxicological impact studies based on Escherichia coli bacteria in ultrafine ZnO nanoparticles colloidal medium. Nano Letters 6:866-870. 6. Brunner, T. J., P. Wick, P. Manser, P. Spohn, R. N. Grass, L. K. Limbach, A. Bruinink, and W. J. Stark. 2006. In vitro cytotoxicity of oxide nanoparticles: Comparison to asbestos, silica, and the effect of particle solubility. Environmental Science & Technology 40:4374-4381. 7. Cai, R., K. Hashimoto, K. Itoh, Y. Kubota, and A. Fujishima. 1991. Photokilling of malignant-cells with ultrafine Ti02 powder. Bulletin of the Chemical Society of Japan 64: 1268-1273. 8. Capone, D. G., D. D. Reese, and R. P. Kiene. 1983. Effects of metals on methanogenesis, sulfate reduction, carbon-dioxide evolution, and microbial biomass in anoxic salt-marsh sediments. Applied and Environmental Microbiology 45: 1 586-1 591. 59 10. ll. 12. 13. 14. 15. 16. 17. 18. Donaldson, K., and P. J. A. Borm. 1998. The quartz hazard: A variable entity. Annals of Occupational Hygiene 42:287-294. Donaldson, K., V. Stone, A. Clouter, L. Renwick, and W. MacNee. 2001. Ultrafine particles. Occupational and Environmental Medicine 58:211-+. Drobne, D. 2007. Nanotoxicology for Safe and Sustainable Nanotechnology. Archives of Industrial Hygiene and Toxicology 58:471-478. Franklin, N. M., N. J. Rogers, S. C. Apte, G. E. Batley, G. E. Gadd, and P. S. Casey. 2007. Comparative toxicity of nanoparticulate ZnO, bulk ZnO, and ZnC12 to a freshwater microalga (Pseudokirchneriella subcapitata): The importance of particle solubility. Environmental Science & Technology 41:8484-8490. F ubini, B., and A. Hubbard. 2003. Reactive oxygen species (ROS) and reactive nitrogen species (RNS) generation by silica in inflammation and fibrosis. Free Radical Biology and Medicine 34: 1507-1516. Goswami, D. Y., D. M. Trivedi, and S. S. Block. 1997. Photocatalytic disinfection of indoor air. Journal of Solar Energy Engineering-Transactions of the Asme 119:92-96. Gupta, A. K., and M. Gupta. 2005. Cytotoxicity suppression and cellular uptake enhancement of surface modified magnetic nanoparticles. Biomaterials 26:1565-1573. Gupta, A. K., and S. Wells. 2004. Surface-modified superparamagnetic nanoparticles for drug delivery: Preparation, characterization, and cytotoxicity studies. Ieee Transactions on Nanobioscience 3:66-73. Heinlaan, M., A. Ivask, I. Blinova, H. C. Dubourguier, and A. Kahru. 2008. Toxicity of nanosized and bulk ZnO, CuO and Ti02 to bacteria Vibrio fischeri and crustaceans Daphnia magna and Thamnocephalus platyurus. Chemosphere 71:1308-1316. Hoet, P. H. M., I. Bruke-Hohlfeld, and O. V. Salata. 2004. Nanoparticles - 6O 19. 20. 21. 22. 23. 24. 25. 26. known and unknown health risks. Journal of Nanobiotechnology 2:12. Hussain, S. M., K. L. Hess, J. M. Gearhart, K. T. Geiss, and J. J. Schlager. 2005. In vitro toxicity of nanoparticles in BRL 3A rat liver cells. Toxicology in Vitro 19:975-983. IARC. 1997. IARC Working Group on the Evaluation of Carcinogenic Risks to Humans: Silica, Some Silicates, Coal Dust and Para-Aramid F ibrils. IARC Monographs on the Evaluation of Carcinogenic Risks to Humans/Would Health 68:1—475. Jin, Y. H., S. Kannan, M. Wu, and J. X. J. Zhao. 2007. Toxicity of luminescent silica nanoparticles to living cells. Chemical Research in Toxicology 20:1126-1133. Karlsson, H. L., P. Cronholm, J. Gustafsson, and L. Moller. 2008. Copper oxide nanoparticles are highly toxic: A comparison between metal oxide nanoparticles and carbon nanotubes. Chemical Research in Toxicology 21: 1726-1732. Lewinski, N., V. Calvin, and R. Drczek. 2008. Cytotoxicity of nanoparticles. Small 4:26-49. Lin, W. S., Y. W. Huang, X. D. Zhou, and Y. F. Ma. 2006. In vitro toxicity of silica nanoparticles in human lung cancer cells. Toxicology and Applied Pharmacology 217:252-259. Long, T. C., N. Saleh, R. D. Tilton, G. V. Lowry, and B. Veronesi. 2006. Titanium dioxide (P25) produces reactive oxygen species in immortalized brain microglia (BV2): Implications for nanoparticle neurotoxicity. Environmental Science & Technology 40:4346-4352. Lyon, D. Y., J. D. F ortncr, C. M. Sayes, V. L. Colvin, and J. B. Hughes. 2005. Bacterial cell association and antimicrobial activity of a C-60 water suspension. Environmental Toxicology and Chemistry 24:2757-2762. 61 27. 28. 29. 30. 31. 32. 33. 34. Maness, P. C., S. Smolinski, D. M. Blake, Z. Huang, E. J. Wolfrum, and W. A. Jacoby. 1999. Bactericidal activity of photocatalytic Ti02 reaction: Toward an understanding of its killing mechanism. Applied and Environmental Microbiology 65:4094-4098. Martinez-Tabche, L., I. G. Cabrera, L. G. Olivan, M. G. Martinez, and C. G. Faz. 2000. Toxic effects of zinc from trout farm sediments on ATP, protein, and hemoglobin concentrations of Limnodrilus hofimeistefi. Journal of Toxicology and Environmental Health-Part a-Current Issues 59:575-583. Matsunaga, T., R. Tomoda, T. Nakajima, N. Nakamura, and T. Komine. 1988. Continuous-sterilization system that uses photosemiconductor powders. Applied and Environmental Microbiology 54: 1330-1333. Nadtochenko, V., N. Denisov, O. Sarkisov, D. Gumy, C. Pulgarin, and J. Kiwi. 2006. Laser kinetic spectroscopy of the interfacial charge transfer between membrane cell walls of E-coli and Ti02. Journal of Photochemistry and Photobiology a-Chemistry { 181 :401 -407. Neal, A. L. 2008. What can be inferred from bacterium-nanoparticle interactions about the potential consequences of environmental exposure to nanoparticles? Ecotoxicology 17:362-371. Nel, A., T. Xia, L. Madler, and N. Li. 2006. Toxic potential of materials at the nanolevel. Science 311 :622-627. Nohynek, G. J., J. Lademann, C. Ribaud, and M. S. Roberts. 2007. Grey goo on the skin? Nanotechnology, cosmetic and sunscreen safety. Critical Reviews in Toxicology 37:251-277. Oberdiirster, G., A. Maynard, K. Donaldson, V. Castranova, J. Fitzpatrick, K. Ausman, J. Carter, B. Karn, W. Kreyling, D. Lai, S. Olin, N. Monteiro-Riviere, D. Warheit, H. Yang, and I. R. F. R. S. I. N. T. S. W. Group. 2005. Principles for characterizing the potential human health effects from exposure to nanomaterials: elements of a screening strategy. Paricle and Fibre Toxicology 2: 1-35. 62 35. 36. 37. 38. 39. 40. 41. 42. 43. Oberdorster, G., E. Oberdorster, and J. Oberdorster. 2005. Nanotoxicology: An emerging discipline evolving from studies of ultrafine particles. Environmental Health Perspectives 113:823-839. Petri—F ink, A., M. Chastellain, L. Juillerat-Jeanneret, A. Ferrari, and H. Hofmann. 2005. Development of functionalized superparamagnetic iron oxide nanoparticles for interaction with human cancer cells. Biomaterials 26:2685-2694. Rimal, B., A. K. Greenberg, and W. N. Rom. 2005. Basic pathogenetic mechanisms in sificosis: current understanding. Current Opinion in Pulmonary Medicine 11:169-173. Rincon, A. G., and C. Pulgarin. 2004. Bactericidal action of illuminated Ti02 on pure Escherichia coli and natural bacterial consortia: post-irradiation events in the dark and assessment of the effective disinfection time. Applied Catalysis B-Environmental 49:99-1 12. Rogers, N. J., N. M. Franklin, S. C. Apte, and G. E. Batley. 2007. The importance of physical and chemical characterization in nanoparticle toxicity studies. Integrated Environmental Assessment and Management 3:303-304. Saito, T., T. Iwase, J. Horie, and T. Morioka. 1992. Mode of photocatalytic bactericidal action of powdered semiconductor Ti02 on mutans Streptococci. Journal of Photochemistry and Photobiology B-Biology 14:369-379. Sawai, J., H. Igarashi, A. Hashimoto, T. Kokugan, and M. Shimizu. 1996. Effect of particle size and heating temperature of ceramic powders on antibacterial activity of their slurries. Journal of Chemical Engineering of Japan 29:251-256. Shi, X. L., V. Castranova, B. Halliwell, and V. Vallyathan. 1998. Reactive oxygen species and silica-induced carcinogenesis. Journal of Toxicology and Environmental Health-Part B-Critical Reviews 1: 1 81-1 97. Sondi, I., and B. Salopek-Sondi. 2004. Silver nanoparticles as antimicrobial agent: a case study on E-coli as a model for Gram-negative bacteria. Journal of Colloid and Interface Science 275:177-182. 63 44. 45. 46. 47. 48. 49. 50. 51. 52. Sunada, K., T. Watanabe, and K. Hashimoto. 2003. Studies on photokilling of bacteria on Ti02 thin film. Journal of Photochemistry and Photobiology a-Chemistry 156:227-233. Swadeshmukul Santra, K. W., Rovelyn Tapec, Weihong Tan. 2001. Development of novel dye-doped silica nanoparticles for biomarker application. Journal of Biomedical Optics 6: 160-166. Thill, A., O. Zeyons, O. Spalla, F. Chauvat, J. Rose, M. Auffan, and A. M. Flank. 2006. Cytotoxicity of Ce02 nanoparticles for Escherichia coli. Physico-chemical insight of the cytotoxicity mechanism. Environmental Science & Technology 40:6151-6156. Vallee, B. L., and D. S. Auld. 1990. Active-site zinc ligands and activated H20 of zinc enzymes. Proceedings of the National Academy of Sciences of the United States of America 87 :220—224. Vega-Lopeza, A., E. Amora-Lazcanob, 1E. Lopez-Lopezc, O. Terrona, and J. B. Proal-Najera. 2007. Toxic effects of zinc on anaerobic microbiota from Zimapan Reservoir (Mexico). Anaerobe 13:65-73. Wei, C., W. Y. Lin, Z. Zainal, N. E. Williams, K. Zhu, A. P. Kruzic, R. L. Smith, and K. Rajeshwar. 1994. Bactericidal activity of Ti02 photocatalyst in aqueous-media - toward a solar-assisted water disinfection system. Environmental Science & Technology 28:934-938. Wei Lian, S. A. L., Hassan Badrane, Weihong Tan, Dongbai Wu, Henry V. Baker, Paul A. Gulig, Daniel V. Lim, and Shouguang Jin. 2004. Ultrasensitive detection of biomolecules with fluorescent dye-doped nanoparticles. Analytical Biochemistry 334: 135-144. Win, K. Y., and S. S. F eng. 2005. Effects of particle size and surface coating on cellular uptake of polymeric nanoparticles for oral delivery of anticancer drugs. Biomaterials 26:2713-2722. Yang, Z. H., and C. S. Xie. 2006. Zn2+ release from zinc and zinc oxide particles in simulated uterine solution. Colloids and Surfaces B-Biointerfaces 47: 140-145. 64 CHAPTER 3 INDUCTION OF INTRACELLULAR REACTIVE OXYGEN SPECIES BY NANOPARTICLES AND THE MEASUREMENT OF GENOTOXICITY BY COMET ASSAY 65 Abstract With the rapid development of nanotechnology, various NMs have been applied in diverse areas, including cosmetics, water purification, material science, electrical devices and drug delivery. The biosafety of NPs has begun to attract attention with regard to their toxicological and environmental effects. I found that zinc oxide (ZnO) and titanium oxide (Ti02) NPs induced reactive oxygen species (ROS) in Escherichia coli strain K12 (Gram-negative) and Exiguobacterium sibiricum strain 255-15 (Gram-positive) in a dose-dependent manner. A 12 h exposure to silicon oxide (Si02) NPs induced ROS in E. sibiricum, but not in E. coli except for 5000 ppm Si02 NPs. There were more ROS induced by Ti02 NPs and Si02 NPs in E. sibiricum than in E. coli, while ZnO NPs induced more ROS in E. coli than in E. sibiricum, which is consistent with our viability results. The 5 nm Ti02 NPs caused more ROS induction than 150 nm Ti02 NPs, which confirmed that size is one of the important factors in NP toxicity. Both Ti02 NPs and ZnO NPs displayed genotoxicity as measured by comet assay, and Ti02 NPs were more genotoxic than ZnO NPs. During 3 h exposure, 10 ppm Ti02 and 1 ppm ZnO NPs caused significant DNA damage in lymphocytes. 66 Introduction The development in NMs and nanotechnology offers various new opportunities in diverse fields, such as electronics, medicines, environmental remediation and food. In 2005, more than $10 billion were spent on the nanotechnology development internationally (21). However, the environmental exposure to the NPs may have some adverse effect on microorganisms, although NPs have existed in the nature for eons (31). Few studies about the influence of NP on public health and environment are available, and there is no clear understanding of the mechanisms of NP toxicity. Recently there have been a few reports on the antimicrobial activity of NPs and interaction between NPs and microorganisms (7, 8, 18, 34, 35, 42). The induction of ROS by NPs is considered as one of the possible mechanisms for NP toxicity. In 1997, Rahman et al. found the increasing induction of ROS by nano-sized Ti02 in human and rat alveolar macrophages (36). Meanwhile, Ti02 NPs were toxic and caused oxidative stress in different cell types, including bacteria and mammal cells (2, 20, 26, 39, 50-52). The induction of ROS by ZnO NPs has also been hypothesized as an explanation for their toxicity, though the mechanisms for their toxicity are not yet well established (3 8). NPs may cause genotoxicity to different cells, since the ROS can enter cells and oxidize membranes, and further damage the DNA, lipids and proteins. There have been a few papers reporting DNA lesions caused by ROS induction from TiOz NPs (16, 20, 24, 27, 50). The alkaline comet assay is a well-established method to detect and measure DNA damage, since it is sensitive, rapid, flexible, cheap and easy to apply (45-47). This 67 technology has been used in different fields, including radiation biology, genotoxicity, clinical research, and genetic ecotoxicology (3, 12, 14). In this study, we investigated the ROS induction by Si02, Ti02 and ZnO NPs in E. coli and E. sibiricum. I also suggested a possible mechanism of ROS induction by Ti02 NPs. The comet assay was performed to evaluate the genotoxicity of Ti02 and ZnO NPs, and Olive tail moment, tail DNA and tail length were measured to quantify DNA damage in lymphocytes. Materials and Methods Reagents and preparation of NP suspensions. RPMI-l640 medium and 2’, 7’- dichlorofluorescin diacetate (DCFH-DA) was obtained from Invitrogen (Carlsbad, CA). Normal melting agarose (NMA), low melting point agarose (LMPA), Histopaque, Phosphate Buttered Saline (PBS), Si02 (10 nm), Ti02 (5 nm and 150 nm) and ZnO (60 nm) were purchased from Sigma-Aldrich (St. Louis, MO). One hundred g/L NPs stock solutions were prepared in distilled and deionized water. Bacterial culture and cell culture. E. coli K12 was grown aerobically in Luria-Bertani (LB) liquid broth at 37 °C and 200 rpm for 12 h. E. sibiricum was cultured in TSB medium at 30 °C and 150 rpm for 12 h. Solid medium for E. coli or E. sibiricum was as above with 1.5% agar (Difco) added. All the cultures with NPS were grown under the same conditions. Intracellular ROS measurement. The intracellular ROS was measured by using 68 DCFH-DA as described previously (25, 49). Ten mM of DCFH-DA stock solution in DMSO was diluted with PBS buffer into 20 M of working solution. E. coli K12 or E. sibiricum was exposed to different concentrations of NPs in 2 ml of liquid media for 24 h. Cells were collected by centrifugation in 1.8 ml of liquid medium, washed twice with water and then incubated in 1.8 ml of DCFH-DA working solution at 37 °C or 30 °C for 1 h. Samples were washed twice with 1X PBS, and their fluorescence was measured at 485-nm excitation and 520-nm emission with SpectraMax M2 spectrometer (Molecular Devices, Sunnyvale, CA). Blank samples (only microorganisms) and control samples (only NPs) were treated in the same way. Viability was measured by plate counting. All the data are the mean of triplicate samples. Alkaline comet assay. Fresh Heparinized venous blood was used for separation of lymphocytes. Blood (100 pl) was diluted with RPMI-164O medium (1 ml) and the dilution was layered over 200 pl Histopaque buffer. After centrifuged at 800>13) for 20 min to allow for unwinding of the DNA, and subsequently electrophoresis was performed at 24 V and 300 mA for 30 min. After that, the slides were washed three times (5 min each) with neutralization buffer (400 mM Tris, pH 7.5) and stained with 1 ug/ml ethidium bromide (80 ul) for 5 min. The slides were then soaked in cold distilled water for 5 min, covered with cover-slips, and then stored at 4 0C in a humidified slide box (no more than 2 days). The observation was made at 400> H202 (3) H202 + 02' —> °OH + OH' + 02 (4) e' + H202 —->°OH + OH' (5) h+ + 011' —+°OH (6) h+ + H20 —+ °OH + H* (7) 2h* + 21120 —> 2H+ + 11202 (8) hv + H202 -> 2'OH (9) An interfacial charge transfer occurred between T102 NPs and the membrane of E. coli (32). The reaction 10 and 11 exhibit the surface charge trapping on T102 NPs. During the process of interfacial charge transfer, h+ reacts with peptidoglycan (PGN) and lipo-polysaccharides (LPS) on microbial cell wall membranes, such as E. coli (reaction 12). As well, h+ can be scavenged by PE, LPS or E. coli during surface oxidation (reaction 13). e' + Ti(IV)O-H —+ Ti(III)O-H' (10) h++Ti(IV)O-H—-Ti (IV)o--H+ (11) h++RH—>R'+H+ (12) Ti(IV)O--H+ + RH —-» OH + R- + H" (13) ROS induction can result in other intracellular damage and disorder of cellular activities. Matsunaga et al. reported that TiOz could oxidize CoA and cause cell death (28, 29). The cell membrane may be oxidized prior to the oxidation of CoA (13, 30, 43). Saito 73 et al. found rapid leakage of K+ ions and slow release of proteins and RNA, and they explained that the cell death was due to a significant disorder in cell permeability (5, 37). Genotoxicity of Ti02 and ZnO NPs. Exposure to Ti02 and ZnO NPs caused significant DNA damage in lymphocytes. Lymphocytes treated with different concentration of T102 NPs for 6 h, and the clear DNA tails, appeared at concentrations greater than 1 ppm. As shown in Table 3.1, the tail length, DNA tail and Olive tail moment increased with the increasing Ti02 NP concentration for both 3 h and 6 h exposures, which meant that DNA damage aggravated as the NP concentration increased. The lymphocytes displayed significant DNA damage in 3 h exposure when the Ti02 NP concentrations were greater than 1 ppm. During the 6 h exposure, the tail length, the tail DNA and the Olive tail moment demonstrated significant difference at concentrations above 100 ppb, compared to the negative control. There is no significant difference between 3 h and 6 h exposures. In the case of ZnO NP treatment, the enhancement of tail length, DNA tail and olive tail moment was demonstrated as the concentration of ZnO NPs increased. The concentrations of ZnO NPs, at which significant DNA damage started, were 10 ppm for 3 h exposure and 1 ppm for 6 h exposure. These results were ten times higher than those for Ti02 NPs. There is no clear explanation about genotoxicity caused by NPs, although one of the hypotheses is ROS induction (8, 16, 19, 20, 35, 40, 43, 48). The results suggest that Ti02 NPs caused more genotoxicity than ZnO NPs, consistent with other reports (23). The explanations are: 74 (a) Dissolution of ZnO NPs. Our previous studies have shown that ZnO dissolves in LB and TSA medium to form Zn2+ ion, which is not genotoxic. Actually low concentrations of Zn2+ ion can stabilize membrane structures and prevent ROS production (6, 41). The saturation concentration of Zn2+ ion in the solution of ZnO NPs is approximately 100-110 ppm. The dissolution of ZnO NPs may reduce the genotoxicity of ZnO NPs. (b) Size difference between T102 NPs and ZnO NPs. The size of Ti02 NPs was 5 nm, while the size of ZnO NPs was 60 nm. For the same mass, Ti02 NPs provide more surface area and hence supply more surface reactive sites than ZnO NPs. There may be more Ti02 NPs absorbed to the lymphocyte membrane than ZnO NPs due to their smaller size. Moreover, it is probably easier for the much smaller Ti02 NPs to enter the membranes of lymphocytes than for the larger ZnO NPs. Induction of ROS has been hypothesized as one of possible mechanisms for genotoxicity. As shown in Figure 3.1, ZnO NPs induced much more ROS than Ti02 NPs. When the concentrations of NPs were lower than 100 ppm, however, there was not much difference between the ROS induced by ZnO NPs and by TiOz NPs, especially in E. sibiricum. Fullerenes can cause DNA damage, and it is much more genotoxic than ZnO NPs and Ti02 NPs (15). The reason may be that fullerenes can disperse more evenly in solution and not aggregate as easily as ZnO NPs and Ti02 NPs. In addition, Silica NPs was reported to caused no genotoxicity by the comet assay (4). In conclusion, the ROS induction by NPs was found in both E. coli and E. sibiricum. 75 SiOz NPs could induce ROS in E. sibiricum when the concentration was greater than 500 ppm, while they did not induce ROS in E. coli except at 5000 ppm. TiOz NPs induced more ROS in E. sibiricum than in E. coli, which is consistent with our previous viability tests. The 5 nm Ti02 NPs induced more ROS than 150 nm Ti02 NPs in either E. sibiricum or E. coli. According to the measurement of Olive tail moment, tail DNA and tail length in comet assay, 1 ppm ZnO NPs in 6 h exposure and 10 ppm ZnO NPs in 3 h exposure caused significant DNA damage in lymphocytes. The 100 ppb Ti02 NPs for 6 h exposure and 1 ppm Ti02 NPs for 3 h exposure damaged DNA significantly, which demonstrated that Ti02 NPs are more genotoxic than ZnO NPs for lymphocytes. 76 ‘0 ‘ - $10, NPS (10 nm) - T10, NPs (5 nm) :1 - Tio, NPs (150 nm) ! I . 8 l E c i: 1 : 1 1 s . i * O - .1 .. 1 . -. “5 r‘ 11 .9 . . '=' “ 2 ii 15 = j 1; '- ~ 1 «1 ’2 i 1 ii 1 . i .‘ 0.1 - control 10 100 500 1000 2000 5000 NPs concentration (ppm) b 100 - - SiO, NPs (10 nm) 1102 NPs (5 nm) 1 - no, NPs (150 nm) I § 1 l 9 1' " s l 1 c :“ E 3. ‘1 i L O 1.1 -- H E 1 “ ‘ g l l 1 1 :1 .3 1.- g 0.1 - control 10 100 500 1000 2000 5000 NPs concentration (ppm) Figure 3.1. (a) The production of ROS by Si02 and Ti02 NPs in E.coli. (b) The production of ROS by $102 and Ti02 NPS in E.sibiricum. 77 1000 * - E. coli m E. sibiricum 100 - 8 C 8 m 9 - i 8 1o - :1 + E .. ' o .2. .1 i i g (B r'w‘ J. a: a i a 1 ‘ 5': .1 a l l 1 0.1 - E3 ’3' ' control 10 50 70 100 200 500 NPs concentration (ppm) Figure 3.1 cont. (c) The production of ROS in E. coli and E. sibiricum by ZnO NPs. 78 .6580 game: 3 nocmano .odva ”modvm sobcoo 0383 Z an EEQEE banEcwi .oZ o a 2.332 $38.2 3.32.? 3.38.3 8.38.8 333% 2558 o o o o o o . . 3.33.0 85H :8 8.35m 3.3. am we; H Rs 3.3 E 02332 .233 2.33; 8.33.2 madam? 2.38.8 $.33: can 2: o o o o o 0 Moon”: waffle; 8; .33 Se Moms 2: name 393%: 6% S o o o o o o 8.3 So Banana ”2 3% gone? 2.3.8.2 8.0.1132 o o o o o o 8% _ we??? 8633 «3 H03. ”Enema 8.. n3: :4 H 3:: . a: 2: n m n a n a emanate 3.3de $4 n mom Seneca of flea a: flaw ea 2 m m m m w a 2.0% :3 8632 2.333 moonmnm 3.13: $3.42.: ea _ m a a m w w a e ._ n a e __ m ._ e __ n .853 =3 2:0 3: <2: .5 as: as. .3 58 3&8 5 $2 No: 3 35. 553 .5502 :3 2:0 .33 <29 :3 .33 532 :32 50505302 .3. 22¢ 79 ._ob=oo gamma: 3 @8888 3.on ”8.on ”.828 03832 Set EBQEE DEScEwE 82 o n m ofimflmfiw— vc.mfloo.m_ owdflmmfim wvfiflwmgm chflmmdc wwdflofimm o>=aom o o o o o o . . 86H mud modflmvd Cdfloofi mmdfi mwd mm; H Ed mm; .IT Saw 03832 86%ng vmdflom; mm; Hmod am; Hwos om... Hms: mm; Homw— Ema oo— o o o o o o mmdfl S; 86% 3d 8.. H 3.0 wwdflofio :1 H Zam— Vm._1_l.t..: Egg 0— o o o o o n wodflnfio omdflncd _v.oflah.m wcdflmné Sdflm—N— mwdeQo— o a o a o a 8% _ vmdflmfio fdflvod oodflnoé mdeo—é C4Hoo.: ozflomd— can oo— m a n m m a 86.4...de vodflnvd nmdflmfim mdem—d wvgfltd 0m._flmo.o can 2 m a a a m a hodflmmd 56H 36 cmdflomd NvdH :.m X: Hmod ovdflpfi £5 _ w m m a a a .— e .— n .— c A n .— e .— n 25:35. :8 3:0 Ae\ev <75 =ah. A55 5»:— =5. 33m 858 5 £2 OcN 3 92:: EEEV .582: :8 «>20 .Ao\ov <20 :9 A83 59.2 :8 go 5082382 .m.m 2an 80 Figure 3.2. Human lymphocytes treated with Ti02 showing genotoxicity (magnification 400): (a, b, c, d, e, and f) nuclei from Ti02-treated 6 h (1 ppb, 10 ppb, 100 ppb, 1 ppm, 10 ppm and 100 ppm) human lymphocytes showing DNA damage; (g) nucleus from an untreated human lymphocyte (negative control); (h) nucleus from an ethyl methanesulfonate (2 mM; positive control) treated human lymphocyte showing DNA damage. 81 References 1. Adams, D. M., L. Brus, C. E. D. Chidsey, S. Creager, C. Creutz, C. R. Kagan, P. V. Kamat, M. Lieberman, S. Lindsay, R. A. Marcus, R. M. Metzger, M. E. Michel-Beyerle, J. R. Miller, M. D. Newton, D. R. Rolison, O. Sankey, K. S. Schanze, J. Yardley, and X. Y. Zhu. 2003. Charge transfer on the nanoscale: Current status. Journal of Physical Chemistry B 107 :6668-6697. 2. Adams, L. K., D. Y. Lyon, A. McIntosh, and P. J. J. Alvarez. 2006. Comparative toxicity of nano-scale Ti02, Si02 and ZnO water suspensions. Water Science and Technology 54:327-334. 3. Bajpayee, M., A. Dhawan, D. Parmar, A. K. Pandey, N. Mathur, and P. K. Seth. 2002. Gender-related differences in basal DNA damage in lymphocytes of a healthy Indian population using the alkaline comet assay. Mutation Research-Genetic Toxicology and Environmental Mutagenesis 520:83-91. 4. Barnes, C. A., A. Elsaesser, J. Arkusz, A. Smok, J. Palus, A. Lesniak, A. Salvati, J. P. Hanrahan, W. H. de Jong, E. Dziubaltowska, M. Stepnik, K. Rydzynski, G. McKerr, I. Lynch, K. A. Dawson, and C. V. Howard. 2008. Reproducible Comet assay of amorphous silica nanoparticles detects no genotoxicity. Nano Letters 8:3069-3074. 5. Bekbolet, M., and C. V. Araz. 1996. Inactivation of Escherichia coli by photocatalytic oxidation. Chemosphere 32:959-965. 6. Bray, T. M., and W. J. Bettger. 1990. The Physiological-Role of Zinc as an Antioxidant. Free Radical Biology and Medicine 8:281-291. 7. Brayner, R., R. Ferrari-Iliou, N. Brivois, S. Djediat, M. F. Benedetti, and F. Fievet. 2006. Toxicological impact studies based on Escherichia coli bacteria in ultrafine ZnO nanoparticles colloidal medium. Nano Letters 6:866-870. 8. Chang, Q. Y., L. Z. Yan, M. X. Chen, H. He, and J. H. Qu. 2007. Bactericidal mechanism of Ag/A1203 against Escherichia coli. Langmuir 23:11197-11199. 82 10. 11. 12. 13. 14. 15. l6. l7. Cho, M., H. Chung, W. Choi, and J. Yoon. 2004. Linear correlation between inactivation of E-coli and OH radical concentration in Ti02 photocatalytic disinfection. Water Research 38: 1069-1077. Cho, M., H. M. Chung, W. Y. Choi, and J. Y. Yoon. 2005. Different inactivation behaviors of MS-2 phage and Escherichia coli in Ti02 photocatalytic disinfection. Applied and Environmental Microbiology 71:270-275. Choi, 0., and Z. Q. Hu. 2008. Size dependent and reactive oxygen species related nanosilver toxicity to nitrifying bacteria. Environmental Science & Technology 42:4583-4588. Cotelle, S., and J. F. Ferard. 1999. Comet assay in genetic ecotoxicology: A review. Environmental and Molecular Mutagenesis 34:246-255. Dadjour, M. F., C. Ogino, S. Matsumura, and N. Shimizu. 2005. Kinetics of disinfection of Escherichia coli by catalytic ultrasonic irradiation with Ti02. Biochemical Engineering Journal 25:243-248. De Boeck, M., D. Lison, and M. Kirsch-Volders. 1998. Evaluation of the in vitro direct and indirect genotoxic effects of cobalt compounds using the alkaline comet assay. Influence of interdonor and interexperimental variability. Carcinogenesis 19:2021-2029. Dbawan, A., J. S. Taurozzi, A. K. Pandey, W. Q. Shan, S. M. Miller, S. A. Hashsham, and V. V. Tarabara. 2006. Stable colloidal dispersions of C-60 fullerenes in water: Evidence for genotoxicity. Environmental Science & Technology 40:7394-7401. Dunford, R., A. Salinaro, L. Z. Cai, N. Serpone, S. Horikoshi, H. Hidaka, and J. Knowland. 1997. Chemical oxidation and DNA damage catalysed by inorganic sunscreen ingredients. Febs Letters 418:87-90. Durand, P. L. O. a. R. E. 2005. Heterogeneity in DNA Damage Using the Comet Assay. Cytometry Part A 66: 1-8. 83 l8. 19. 20. 21. 22. 23. 24. 25. 26. 27. Fu, G. F., P. S. Vary, and C. T. Lin. 2005. Anatase Ti02 nanocomposites for antimicrobial coatings. Journal of Physical Chemistry B 109:8889-8898. Gogniat, G., M. Thyssen, M. Denis, C. Pulgarin, and S. Dukan. 2006. The bactericidal effect of Ti02 photocatalysis involves adsorption onto catalyst and the loss of membrane integrity. F ems Microbiology Letters 258: 1 8-24. Gurr, J. R., A. S. S. Wang, C. H. Chen, and K. Y. Jan. 2005. Ultrafine titanium dioxide particles in the absence of photoactivation can induce oxidative damage to human bronchial epithelial cells. Toxicology 213:66-73. Harrison, P. 2007. Emerging challenges: nanotechnology and the environment. GEO Year book 2007:61-68. Kang, S., M. Pinault, L. D. Pfefferle, and M. Elimelech. 2007. Single-walled carbon nanotubes exhibit strong antimicrobial activity. Langmuir 23:8670-8673. Karlsson, H. L., P. Cronholm, J. Gustafsson, and L. Moller. 2008. Copper oxide nanoparticles are highly toxic: A comparison between metal oxide nanoparticles and carbon nanotubes. Chemical Research in Toxicology 21:1726-1732. Kong, S. J., B. M. Kim, Y. J. Lee, and H. W. Chung. 2008. Titanium dioxide nanoparticles trigger p53-mediated damage response in peripheral blood lymphocytes. Environmental and Molecular Mutagenesis 49:399-405. Lin, W. S., Y. W. Huang, X. D. Zhou, and Y. F. Ma. 2006. In vitro toxicity of silica nanoparticles in human lung cancer cells. Toxicology and Applied Pharmacology 217 :252-259. Long, T. C., J. Tajuba, P. Sama, N. Saleb, C. Swartz, J. Parker, S. Hester, G. V. Lowry, and B. Veronesi. 2007. Nanosize titanium dioxide stimulates reactive oxygen species in brain microglia and damages neurons in vitro. Environmental Health Perspectives 115: 1631-1637. Mates, J. M., and F. M. Sanchez-Jimenez. 2000. Role of reactive oxygen 84 28. 29. 30. 31. 32. 33. 34. 35. species in apoptosis: implications for cancer therapy. International Journal of Biochemistry & Cell Biology 32: 157-170. Matsunaga, T., R. Tomoda, T. Nakajima, N. Nakamura, and T. Komine. 1988. Continuous-sterilization system that uses photosemiconductor powders. Applied and Environmental Microbiology 54: 1330-1333. Matsunaga, T., R. Tomoda, T. Nakajima, and H. Wake. 1985. Photoelectrochemical sterilization of microbial-cells by semiconductor powders. F ems Microbiology Letters 29:211-214. Mitoraj, D., A. Janczyk, M. Strus, H. Kiscb, G. Stochel, P. B. Heczko, and W. Macyk. 2007. Visible light inactivation of bacteria and fungi by modified titanium dioxide. Photochemical & Photobiological Sciences 6:642-648. Murr, L. E., E. V. Esquivel, J. J. Bang, G. de la Rosa, and J. L. Gardea-Torresdey. 2004. Chemistry and nanoparticulate compositions of a 10,000 year-old ice core melt water. Water Research 38:4282-4296. Nadtochenko, V., N. Denisov, O. Sarkisov, D. Gumy, C. Pulgarin, and J. Kiwi. 2006. Laser kinetic spectroscopy of the interfacial charge transfer between membrane cell walls of E-coli and Ti02. Journal of Photochemistry and Photobiology a-Chemistry 181:401-407. Nair, S., A. Sasidharan, V. C. Divya Rani, D. Menon, K. M. Seema Nair, and S. Raina. 2008. Role of size scale of ZnO nanoparticles and microparticles on toxicity toward bacteria and osteoblast cancer cells. Journal of Materials Science. Materials in Medicine. Oh, S. D., B. S. Byun, S. Lee, S. H. Choi, M. 1. Kim, and H. G. Park. 2007. Radiolytic synthesis of ag-loaded polystyrene (Ag-PS) nanoparticles and their antimicrobial efficiency against Staphylococcus aureus and Klebsiella pneumoniase. Macromolecular Research 15:285-290. Pal, S., Y. K. Tak, and J. M. Song. 2007. Does the antibacterial activity of silver nanoparticles depend on the shape of the nanoparticle? A study of the gram-negative bacterium Escherichia coli. Applied and Environmental 85 36. 37. 38. 39. 40. 41. 42. 43. 44. Microbiology 73:1712-1720. Rahman, Q., J. Norwood, and G Hatch. 1997. Evidence that exposure of particulate air pollutants to human and rat alveolar macrophages leads to differential oxidative response. Biochemical and Biophysical Research Communications 240:669-672. Saito, T., T. Iwase, J. Horie, and T. Morioka. 1992. Mode of photocatalytic bactericidal action of powdered semiconductor Ti02 on Mutans Streptococci. Journal of Photochemistry and Photobiology B-Biology 14:369-379. Sawai, J., E. Kawada, F. Kanou, H. Igarashi, A. Hashimoto, T. Kokugan, and M. Shimizu. 1996. Detection of active oxygen generated from ceramic powders having antibacterial activity. Journal of Chemical Engineering of Japan 29:627-633. Sayes, C. M., R. Wahi, P. A. Kurian, Y. P. Liu, J. L. West, K. D. Ausman, D. B. Warheit, and V. L. Colvin. 2006. Correlating nanoscale titania structure with toxicity: A cytotoxicity and inflammatory response study with human dermal fibroblasts and human lung epithelial cells. Toxicological Sciences 92:174-185. Sondi, 1., and B. Salopek-Sondi. 2004. Silver nanoparticles as antimicrobial agent: a case study on E-coli as a model for Gram-negative bacteria. Journal of Colloid and Interface Science 275:177-182. Stobs, S. J., and D. Bagchi. 1995. Oxidative mechanisms in the toxicity of metal-ions. Free Radical Biology and Medicine 18:321-336. Stoimenov, P. K., R. L. Klinger, G. L. Marchin, and K. J. Klabunde. 2002. Metal oxide nanoparticles as bactericidal agents. Langmuir 18:6679-6686. Sunada, K., T. Watanabe, and K. Hashimoto. 2003. Studies on photokilling of bacteria on Ti02 thin film. Journal of Photochemistry and Photobiology a-Chemistry 156:227-233. Thill, A., O. Zeyons, O. Spalla, F. Chauvat, J. Rose, M. Auffan, and A. M. 86 45. 46. 47. 48. 49. 50. 51. 52. Flank. 2006. Cytotoxicity of Ce02 nanoparticles for Escherichia coli. Physico-chemical insight of the cytotoxicity mechanism. Environmental Science & Technology 40:6151-6156. Tice, R. R. 1995. The single cell gel/comet assay: A microgel electrophoretic technique for the detection of DNA damage and repair in individual cells. Environmental mutagenesis:315-339. Tice, R. R., E. Agurell, D. Anderson, B. Burlinson, A. Hartmann, H. Kobayashi, Y. Miyamae, E. Rojas, J. C. Ryu, and Y. F. Sasaki. 2000. Single cell gel/comet assay: Guidelines for in vitro and in vivo genetic toxicology testing. Environmental and Molecular Mutagenesis 35:206-221. Tice, R. R., P. W. Andrews, 0. Hirai, and N. P. Singh. 1991. The single cell gel (SCG) assay: an electrophoretic technique for the detection of DNA damage in individual cells. 283: 157—164. Wamer, W. G., J. J. Yin, and R. R. Wei. 1997. Oxidative damage to nucleic acids photosensitized by titanium dioxide. Free Radical Biology and Medicine 23:851-858. ’ Wang, H., and J. A. Joseph. 1999. Quantifying cellular oxidative stress by dichlorofluorescein assay using microplate reader. Free Radical Biology and Medicine 27 :612-616. Wang, J. J., B. J. S. Sanderson, and H. Wang. 2007. Cyto- and genotoxicity of ultrafine Ti02 particles in cultured human lymphoblastoid cells. Mutation Research-Genetic Toxicology and Environmental Mutagenesis 628:99-106. Warheit, D. B., T. R. Webb, K. L. Reed, S. Frerichs, and C. M. Sayes. 2007. Pulmonary toxicity study in rats with three forms of ultrafine-Ti02 particles: Differential responses related to surface properties. Toxicology 230:90-104. Warheit, D. B., T. R. Webb, C. M. Sayes, V. L. Colvin, and K. L. Reed. 2006. Pulmonary instillation studies with nanoscale Ti02 rods and dots in rats: Toxicity is not dependent upon particle size and surface area. Toxicological Sciences 91:227-236. 87 53. Zhang, L. Z., J. C. Yu, H. Y. Yip, Q. Li, K. W. Kwong, A. W. Xu, and P. K. Wong. 2003. Ambient light reduction strategy to synthesize silver nanoparticles and silver-coated Ti02 with enhanced photocatalytic and bactericidal activities. Langmuir 19: 10372-10380. 88 CHAPTER 4 THE INFLUENCES OF OXIDE NANOPARTICLES ON GENE EXPRESSION OF ESCHERICHLA COL! K12 89 Abstract Si02, Ti02, and ZnO NPs are used in various fields as support or catalysts. Our previous toxicity data demonstrated their antibacterial activities. In this study, we explored their influence on the gene expression of E. coli, which generally increased from Si02 to Ti02 to ZnO NPs. In Si02 and ZnO NPs treatments, the number of repressed genes was higher than that of the induced ones, while more genes were induced than the repressed in Ti02 exposure. The transporters for amino acid and carbohydrate were down-regulated, while the transporter for iron or a few other metal ions were up-regulated. Moreover, stress-related genes and DNA repair genes were generally induced. This may be due to a number of reasons including ROS induction, damage to membrane, proteins or DNA, alteration of membrane permeability and fluidity, attack to iron-sulfur cluster and nick on DNA. It was also demonstrated that ZnO NPs influenced microbial respiration. The microarray data demonstrated that NPs regulated the genes in different categories, and E. coli adjusted their metabolic pathways and defense mechanisms during the exposure to NPS. 90 Introduction Nanotechnology, defined by The Royal Society, is “the design, characterization, production and application of structures, devices and systems by controlling shape and size at nanometer scale” (29). The growth in nanotechnology offers numerous technological advances and industrial development opportunities, and it is estimated that the global market for nanotechnology grow to about $1 trillion by 2015. NMs, one proportion of nanotechnology, possess enhanced and unique mechanical, optical and electrical characteristics, which broaden their application in diverse fields, such as environmental remediation, cosmetics, medicine and catalysis. Although the number of studies addressing to the toxicity of NPs is increasing, there are many unknowns about the environmental impacts of NPs and the mechanisms of NP toxicity (70, 81). The toxicity of NPs to microorganisms is an important concern for evaluating the environmental impact of NPs. Microorganisms play a fundamental role in the environment, and supply key environmental services in maintaining healthy ecosystems. Studying toxicity of NPs on microorganisms is beneficial for us to evaluate the environmental impacts of NPs and explore the nanotoxicology on animals and human. Possible explanations for NP toxicity include induction of ROS, disruption of membrane, releases of hazardous constituents, and damage to DNA (52). $02, Ti02, and ZnO NPs are popular NPs, and used for cosmetics, food, painting and electrical devices. Several papers have been published about their antibacterial activities (3, 7, 32, 33, 80). In this study, E. coli K12 is chosen as the test species to 91 explore gene expression influenced by Si02, Ti02, and ZnO NPs. E. coli K12, Gram-negative, facultative anaerobic and non-spore forming, is one of the most studied bacteria in the field of microbiology. Microarray technique was applied to explore the gene expression of E. coli in the NP exposure. Microarray data were validated by quantitative PCR for a few target genes. We categorized the influenced genes, and also analyzed and explored the pathways and the transcriptional units that are regulated by NPs. Materials and Methods Reagents and preparation of NP suspensions. Random hexamer primers, 100 mM nucleotide set, RNaseOUT, salmon sperm DNA, superscriptase 11 RT, formamide, 10% sodium dodecyl sulfate, 20x Saline-Sodium Citrate (SSC), and 0.5 M ethylenediaminetetraacetic acid (EDTA) were obtained from Invitrogen Co. (Carlsbad, CA). Bovine serum albumin (BSA), dimethyl sulfoxide, sodium hydroxide, 1 M hydrochloric acid and isopropanol were purchased from Sigma-Aldrich Chemical Co. (St. Louis, MO). RNAlater, SlideHyb buffer and aminoallyl—modified dUTP were obtained from Ambion Inc. (Austin, TX). L'ysozyme, DNase I, Tris-Cl EDTA (TE) buffer, RNeasy mini kit, Quiaquick PCR purification kit and RNA protect bacteria reagents were purchased from Qiagen Co (Valencia, CA). Si02 (10 nm), Ti02 (5 nm) and ZnO (60 nm) were purchased from Sigma-Aldrich (St. Louis, MO). One hundred g/L NPs stock solutions were prepared in deionized and distilled water. Phosphate wash buffer (100 ml) 92 for cDNA purification was made of 0.5 m1 K2HPO4 (1 M, pH 8.5), 84.25 ml ethanol and 15.25 ml H20 and phosphate elution buffer (100 ml) contains 0.4 ml KzHPO4 (l M, pH 8.5) and 99.6 ml H20. All solutions were autoclaved and stored at room temperature Cell growth, harvest and RNA extraction. From the freezer stock of E. coli streak plates were prepared with LB medium and grown at 37 °C overnight. One colony on the plate was transferred to the tubes with 5 ml LB broth. The NP stock solution was added to each tube with 10 ml LB broth to result in final NP concentrations of 1000 ppm for Si02 NPs, 1000 ppm for Ti02 NPs and 50 ppm for ZnO NPs. The samples were inoculated and incubated until mid-log growth phase of E. coli (O.58000x g. DNase digestion was applied to the column by buffer RWl washing, DNase I solution washing (10 pl) and incubating, and then buffer RWl re-washing. The RNA was purified as per the protocol of RNeasy mini kit and the 93 final concentration of RNA was measured by Nanodrop from Therrno Scientific Co. (Wilmington, DE). RNA from control samples was processed in a similar way as NP samples. All experimental and control samples were prepared in triplicate. cDNA synthesis and dye-labeling. For each sample, 5 pg of total RNA was used to synthesize cDNA and two RNA populations of each sample were needed for dye-swapping. RNA mixture contained 5 pg RNA, 5 p1 random hexamer primers (3 mM), Rnase-free water in the final volume of 16.8 pl. The RNA mixture was incubated at 70°C for 10 min and then cooled on ice for another 10 min. The RT-PCR reaction had a 30 pl volume and contained 6 pl 5x first strand synthesis buffer, 3 pl DTT (0.1 M), 1 pl RNaseOUT, 1.2 pl 25x dNTPs (7.5 mM aa-dUTP, 5 mM dTTP, 12.5 mM dCTP, 12.5 mM dGTP and 12.5 mM dATP), 2 pl Superscript 11 RT and 16.8 pl RNA mixture. The RT-PCR reaction was incubated at 42°C overnight, and then 10 p1 NaOH (1 M) and 10 pl EDTA (0.5 M) were added to the reaction. After incubating at 65°C for 15 min, 10 pl HCl was added to neutralize the pH. The cDNA was purified according to the protocol of Qiagen purification kit with someimodification. The cDNA mixture was mixed with 300 pl PB buffer, and transferred to a Quiaquick column. After 1 min centrifuging at 17900x g, the collection tube was emptied and the column was washed with 750 pl phosphate wash buffer twice and centrifuged for 1 min. After 1 min additional centrifuging, the column was transferred to a new tube, and then eluted twice by incubating the column with 30 pl phosphate elution buffer for 1 min and centrifuging for 1 min at maximum speed. The cDNA sample was dried using speedvac from Global Medical Instrumentation 94 Inc. (Ramsey, MN). The aa-dUTP cDNA was resuspended in 4.5 p1 sodium carbonate (0.1 M, pH 9.0) and mixed with 4.5 p1 Cy5 or Cy3. After incubating 2 h at room temperature in the dark, 35 pl NaOAc (100 mM, pH 5.2) was added and the dye-coupled cDNA' was purified as described in the protocol of Qiaquick purification kit. The concentrations of cDNA and dye were measured by Nanodrop from Thermo Scientific Co. (Wilmington, DE). Hybridization. The glass-slide arrays were cross-linked at 6000 pJ UV and water bath was pre-warrned to 42°C. The slides were placed in pre-hybridization buffer containing 5x SSC, 0.1% SDS and 0.1 mg/ml BSA at 42°C for 90 min. After washed twice with 0.1x SSC for 5 min and twice with water for 0.5 min, the slides were dried by centrifuging at 1600 rpm for 3 min and transferred to dry and clean tubes covered with foil. The cover-slips were washed with 2% SDS, water and ethanol, followed by additional washing twice with water. Then, the cover-slips were dried with filtered pressured air and placed in clean tubes covered with foil. The labeled DNA of the control sample was re-suspended in 4 pl EDTA, and heated at 95°C for 30 s, and then spun down. The other differently labeled DNA from NP-treated sample was suspended in the same 4 pl EDTA, which was heated at 95°C for 5 min and spun down. The dry slide-array was placed in a clean and dry hybridization chamber and coverslip was placed on the slide. The 40 p1 SlideI—lyb buffer, which was preheated to 68°C, was added to the sample and mixed by pipetting (preventing bubbles). All the hybridization solution was loaded on the slide, and 25 pl water was added in each well in the hybridization chamber. The chamber 95 was placed in water-bath at 42°C for 18 h. For each NP treatment, six hybridizations from three biological replicates and two technical replicates (dye-swapping) were performed. Washing, scanning and data analysis. After 18 h, the chamber was removed from waterbath and dried with paper towel before open. Wash Buffer 1, Wash Buffer 2 and slide staining jars were prewarmed at 37 °C. The slide was taken out with forceps, washed once with Wash Buffer 1 (2x SSC and 0.5% SDS) to remove the coverslip on belly dancer for 5 min, then twice with Wash Buffer 2 (0.1x SSC and 0.1% SDS) on belly dancer for 5 min, and twice with Wash Buffer 3 (0.1x SSC) for 2.5 min. The slide was dried by centrifuging at 1800 rpm for 3 min and transferred to dry tubes. Axon Genepix 4000B laser scanner was used for slide scanning and Genepix 3.0 software was used to produce a raw data set. The scanner was turned on and warmed up for 15 min while the computer was off. The slide was placed face-down, and PMT and power % were adjusted to optimal by verifying histogram. The resolution to 5 pixels per pm and seaming area was selected. Then the interested area was scanned and the image was saved. GENESPRING 6.0 software (Silicon Genetics, Redwood City, CA) was used for data analysis. The data was normalized both per chip and per gene (Lowess method), and dye-swapping. Only the spots with more than 55% of pixel greater than background plus 2x standard deviation in either Cy5 or Cy3 channel were used for analysis. Genes that showed a statistically significant change in gene expression (p>0.05) and a greater than or equal to 2-fold change in magnitude were regarded as significant. Validation of micoarray data by quantitative PCR. For control samples and 96 ZnO treated samples, 5 pg of total RNA was used to synthesize cDNA. RNA mixture contained 5 pgRNA, 5 pl random hexamer primers (3 mM), Rnase-free water in the final volume of 16.8 pl. The RNA mixture was incubated at 70°C for 10 min and then cooled on ice for another 10 min. The reverse transcriptase PCR reaction had a 30 pl volume and contained 6 pl 5x first strand synthesis buffer, 3 pl DTT (0.1 M), 1 pl RNaseOUT, 1.2 pl 25x dNTPs (12.5 mM dTTP, 12.5 mM dCTP, 12.5 mM dGTP and 12.5 mM dATP), 2 pl Superscript 11 RT and 16.8 pl RNA mixture. The RT-PCR reaction was incubated at 42°C overnight, and then 10 pl NaOH (1 M) and 10 pl EDTA (0.5 M) were added to the reaction. After incubating at 65 °C for 15 min, ten pl HCl was added to neutralize the pH. cDNA was purified according to the protocol of Qiagen purification kit with some modification. The cDNA mixture was mixed with 300 pl PB buffer, and transferred to a Quiaquick column. After 1 min centrifuging at 17900x g, the collection tube was emptied and the column was washed by 750 pl phosphate wash buffer twice and centrifuged for 1 min as above. After 1 min additional centrifuging, the column was transferred to a new tube, and then eluted twice by incubating the column with 30pl phosphate elution buffer for 1 min and centrifuging for 1 min at maximum speed. For quantitative PCR, the forward primers and the reverse primers were designed for ten target genes and one reference gene (Table 4.1). The real time PCR reaction had a final volume of 15 pl and contained 7.5 pl PCR green master mix, 0.9 p1 primer mixture (5 pM for forward primer and reverse primer respectively), 5.6 pl PCR water and 1 pl DNA template. Real time PCR was performed on ABI Prism 7900HT Sequence Detection System and the cycling 97 parameters were 50 °C for 2 min, 95 °C for 10 min, and 40 cycles of 95 °C for 15 sec and 60 °C for 1 min. The dissociation curves were run for product identification. Real time PCR was done on ABI Prism 7900HT Sequence Detection System from Applied Biosystems (Foster City, CA). Comparative CT Method (AACT) was applied for quantitative PCR analysis. The gene expression fold is equal to 2AACT and AACFACT test sample —ACT control sample. Results Global gene expression profiles under NP exposure. In the treatments of Si02 and ZnO NPs, a greater number (279 for ZnO and 52 for Si02) of genes were repressed than the induced (197 for ZnO and 2 for Si02). For Ti02 NPs, there were more up-regulated genes (28 genes) compared to the number of repressed genes (11 genes). The genetic response to 100 ppm ZnO NPs was with more than 8.0% of the genome expressing differentially whereas only approximately 1% of genome showed differential expression for both 1000 ppm Ti02 NPs and 1000 ppm Si02 NPs respectively. In our previous studies, E. coli K12 was more tolerant to TiOz and SiOz NPs, but more vulnerable to ZnO NPs, compared to E. sibiricum. During the 1000 ppm Ti02 NPs exposure, there were 289 down-regulated genes and 197 up-regulated genes in E. sibiricum, while 1] induced genes and 28 repressed genes were found in E. coli at the same concentration of Ti02 NPs. Not many genes were common among these three different treatments. According to 98 the l way-ANOVA analysis, there were 499 genes expressing differently in these three NP treatments. Gene expression profile following ZnO treatment. In the ZnO treatment, globally the expression fold of 5,969 genes ranges from 0.02 to 12.8], while there were 279 down-regulated genes and 197 up-regulate genes. In the treatment of ZnO NPs, there were more repressed and induced genes than in the treatment of Si02 or Ti02 NPs. The up-regulated genes in ZnO treatment. The categories “transportation” (19.7%), “transcription” (12.6%) and “translation” (10.1%) were the biggest ones for the up-regulated genes in ZnO treatment. The transporters genes included those for carbohydrate, amino acid, ion, proton, and heme, such as gene glpE rha]? tsp, and nikD (9, 79, 91, 97). Induction of some transporter genes was the response to Zn2+ in ZnO NP suspension. For instance, the induced gene zntA (3.8 fold), coding zinc, cobalt and lead efflux system, was found to be activated by an increased concentration of Zn(II) and Cd(II), showing greater induction by Cd than Zn (85). Gene znuA, coding subunit of ABC transporter, was also induced by Zn2+ ions (2.1 fold) (100). Moreover, gene erA (3.5 fold), coding glucitol/sorbitol-specific enzyme IIC component of PTS, was up-regulated, and its expression was reported to be elevated under basic conditions and addition of Zn to culture grown in minimal medium (58, 65, 87). The aerobic/anaerobic conditions in ZnO NP suspension also affected the expression of a few transporter genes. For example, gene ulaA, coding L-ascorbate transporting phosphotransferase system, was up-regulated 99 (2.6 fold), which is required for the ability to utilize L-ascorbate as the carbon source under anaerobic growth conditions (105). It was also found that induced gene nark (4.6 fold), coding Nark MFS nitrate/nitrite antiporter in anaerobic nitrate respiration, was regulated by nitrate and oxygen (53). Moreover, a threonine sugar transport protein (STP) transporter, coded by gene tch, was induced (2.1 fold), and its expression was reported under anaerobic conditions (86). The aerobic/anaerobic conditions in ZnO suspension also influenced the expression of some other genes, especially for those involved in both anaerobic and aerobic respiration. Both cyoB and cyoD were up-regulated (2.6 and 2.1 fold respectively), coding subunits of cytochrome b0 terminal oxidase in aerobic respiration. The expression of gene aldA was also induced (3.0 fold), which codes subunit of aldehyde dehydrogenase A. Aldehyde dehydrogenase A was expected to be only present under aerobic conditions and be regulated by catabolite repression, because previous work showed that it was repressed under anaerobic conditions and induced by the carbon source (30, 75). Meanwhile, gene sdhC was also up-regulated (2.4 fold), which codes succinate dehydrogenase membrane protein in aerobic respiration, as a large subunit of cytochrome b556. Gene ndh is another example, and it codes NADqubiquinone oxidoreductase II in aerobic respiratory chain (2.0 fold) (69, 73). Genes involved in anaerobic respiration were also induced, such as genes narG and narH (2.3 and 2.1 fold), coding subunits of nitrate reductase A in anaerobic respiration. Nitrate reductase A functions as an electron acceptor and benefits proton motive force (11, 93). These 100 suggested that both anaerobic and aerobic respiration is important for survival in ZnO NP solution. In the category “transcription”, Induction of some genes were due to the stress caused by ZnO NPs. TreR transcriptional repressor, coded by gene treR, was induced by 2.4 fold, and treR can de-repress operons related to carbohydrate transport and catabolism under osmotic stress (44). The induced gene yiiA (2.9 fold), coding CpxR transcriptional dual regulator, can sense a variety of stress and activate some stress-combative genes, such as those involved in drug efflux (42). In the presence of zinc, gene zraS (2.5 fold) was induced, responding to environmental signals. Meanwhile, gene asr (3.3 fold) was also up-regulated, which responds to zinc and plays a role in survival under acid conditions (58, 84). Furthermore, induced gene thA (2.2 fold) codes protein involved in stress resistance, as well as phoH, responding to stress, such as starvation (2.1 fold) (28, 65, 104). All of these demonstrated that E. coli responded to stress resulting from ZnO NP by inducing some gene expression. Genes about cell structure and division were also induced. For example, gene ydaD (2.3 fold) implicated in cell division and cycling was up-regulated, whose over-expression inhibits cell division and leads to morphological defects (21). Gene ych was also induced, which involves in flagellar synthesis, swarming, and cell elongation (26). Moreover, induced gene mukB, coding cell division protein, involves in chromosome segregation and condensation, and also fimctions in cell division and cycle. The induction of these genes is probably due to the slow growing rate of bacteria, caused 101 by ZnO NPs. The categories about metabolism and biosynthesis of carbohydrate, lipid, amino acid or nucleotide were also important in the ZnO NP treatment. For instance, gene sucB and sucD were up-regulated, functioning in tricarboxylic acid cycle. The changes in these categories are probably due to the ROS induction by ZnO NPs, since ROS can oxidize amino acid and DNA, and also interrupt nutrient uptake. Since ZnO NPs can cause genotoxicity, the genes related to DNA repair and RNA modification were found to be up-regulated. Up-regulation of gene dpiA can induce some genes involving citrate fermentation and SOS response, or repress some transcriptional regulator. The subunit of DNA gyrase, coded by induced gene gyrB, involves in DNA modification and replication, as well as transcription. Meanwhile, gene recC, involved in DNA repair and essential for recombination and conjugation as well as repair of double-strand breaks, was also up-regulated (22). Moreover, gene mch which codes for subunit of S-methylcytosine restriction system and defends cell against foreign DNA, was induced. Genes involving in detoxification was also up-regulated. For instance, induction of gene dsdA, coding D-serine deaminase, serves in a catabolic function and detoxification, since D-serine is still a good nitrogen source and moderate carbon source. Moreover, gene frmA, coding formaldehyde dehydrogenase, was also induced, which can detoxify exogenous formaldehyde and be induced by formaldehyde (3 8, 40). Not only single genes, quite a few operons in different categories was also found to be induced by ZnO NPs, for example, the transcriptional unit rplNXE-rpsNH—rpIFR 102 -rpsE-rme-rplO-sec Y -rpmJ and rsz-rplCD WB-rpsS-rplV-rpsC-rpIP-rme-rpsQ. These two big operons encode a serial of SOS and 308 ribosomal subunit proteins. The transcription unit rpsLG-fusA-tqu is another induced operon in the category “translation”, containing genes tqu (elongation factor Tu), fusA (elongation factor G) and rpsG (3OS ribosomal subunit protein S7), and it was estimated that elongation factor Tu was the most abundant protein in E. coli and maximally expressed during stress response (68). Furthermore, both of the transcriptional unit ngABC and glpT Q were induced, coding glycerol-3-phosphate dehydrogenase under anaerobic conditions and glycerol-3-P MFS transporter respectively, which are important for utilization of glycerol (99). The intracellular level of glycerol-3-phosphate is important for biosynthesis of phospholipids. Genes narG (a subunit of nitrate reductase A) and narH ([3 subunit of nitrate reductase A), in the operon narGHJI, was up-regulated, constituting catalytic dimmer of nitrate reductase A in anaerobic respiration. A few operons in the category “energy production” were also induced by ZnO NPs. In the transcriptional unit fdnGHI, fdnG (a subunit of formate dehydrogenase N), fdnH (B subunit of formate dehydrogenase N) and fan (7 subunit of formate dehydrogenase N) exhibited induction, and their expression have been found to be induced by nitrate and anaerobiosis (35, 92). The formate dehydrogenase N is beneficial to generate the proton motive force and allow generation of the ubiquitous energy carrier ATP by ATP synthase. One ATP synthase, coded by the transcription unit athEFHAGDC, was induced, proved 103 by induction of gene atpC (8 subunit of ATP synthase), atpD ([3 subunit of ATP synthase), ath (7 subunit of ATP synthase), atpA (a subunit of ATP synthase), atpH (5 subunit of ATP synthase), and ath (b subunit of ATP synthase), and this ATP synthase can transport proton and synthesize energy. Moreover, in the transcriptional unit napFDAGHBC—ccmABCDEFGH, ccmA (subunits of CcmABCDE protoheme IX ABC transporter), ccmB (subunits of CcmABCDE protoheme IX ABC transporter), and ccmG (CcmEFGH holocytochrome c synthetase), and napC (cytochrome c protein), napH (ferredoxin-type protein), napG (ferredoxin-type protein), and napA (large subunit of periplasmic nitrate reductase), were induced, which were required for electron transfer from ubiquinol (12, 13, 31). Many genes transporting proton were up-regulated during ZnO exposure, involved in maintaining the proton motive force and generate energy. Down-regulated gene profile in the exposure to ZnO NPs. The categories “transportation” (19.0%) and “hypothetical proteins” (10.8%) had the highest fraction of down-regulated genes. For instance, the gene manX, manY and manZ in category “transportation” were repressed, which construct mannose PTS permease and transport exogenous hexoses (76). The transcriptional unit gatYZABCD was also down-regulated, coding tagatose-l ,6-biphosphate aldolase (GatYZ), galactitol-l-l phosphate dehydrogenase (GatD), and galactitol PTS permease (GatABC) that take up and transport exogenous galactitol. A few genes related to the metabolism of carbohydrate and amino acid (totally 11.1%) were down-regulated too, maybe due to the 104 hindrance of their uptake. Moreover, other categories of genes that were down-regulated included “transcription” (6.8%), “response to stress” (5.0%), “cell division and motion” (3.9%), “nucleotide biosynthesis” (3.6%), “DNA repair” (2.9%), RNA modification (2.9%) and “signal transduction” (2.9%). The genes related to anaerobic respiration were also found to be down-regulated. On one hand, some of genes were repressed by the presence of oxygen. For example, the transcription unit hyaABCDEF was repressed, which codes hydrogenase 1. This enzyme consists of HyaA (small subunit of hydrogenase 1), HyaB (large subunit of hydrogenase 1), HyaC (b-type cytochrome subunit of hydrogenase 1), HyaD (protein involved in processing of HyaA and HyaB proteins), HyaE (protein involved in processing of HyaA and HyaB proteins), HyaF (protein involved in nickel incorporation into hydrogenase 1 proteins), and it is tolerant to oxygen and provide complementary redox properties to the cell (57). The hya operon was found to be induced under anaerobic conditions and by the presence of formate, or induced under acidic conditions, but repressed by nitrate (51, 78). Meanwhile, gene fa’hF, coding formate dehydrogenase, was also down-regulated,'and it was found to be repressed by nitrate, nitrite and the presence of oxygen (1). On the other hand, our results also displayed the repression of genes that was reported in anaerobic conditions. For instance, the repressed gene aldA codes NAD-linked aldehyde dehydrogenase A, and its repression occurs under anaerobic conditions via ArcA (59). Under anaerobic conditions, protein ArcA also mediates the repression of L-lactate 105 dehydrogenase, coded by gene lldD, which functions in aerobic respiration and anaerobic nitrate respiration (47). Repressed gene poxB codes pyruvae oxidase, which is responsible for pyruvate metabolism under aerobic conditions, and its expression decreases under anaerobic conditions (2, 20). Aerobic/anaerobic conditions in ZnO NP suspension exhibited down-regulation of some operons. In the transcriptional unit ynfEFGH-dmsD, gene ynfE (oxidoreductase subunit paralog of DmsA), yan (oxidoreductase subunit), ynfG (oxidoreductase, Fe-S subunit paralog of DmsB), yan (DMSO reductase anchor subunit, paralog of DmsC) and dmsD (DMS reductase maturation protein) were repressed, which can utilize dimethyl sulfoxide as a terminal oxidant under anaerobic conditions. The down-regulation of some transcriptional units was also considered as response to the stress caused by ZnO NPs. The transcriptional unit uxaCA, including uxaA (D-altronate dehydratase) and uxaC (D-glucuronate isomerase/D-galacturonate isomerase) was repressed. It was reported that Zn2+ ion could inhibit the enzyme (8). In the transcriptional unit hdeAB-yhiD, hdeA (inactive form of acid-resistance protein), hdeB (acid stress chaperone), yhiD (Mg2+ transport ATPase) were also down-regulated, and all of them involved acid resistance, as well as gene slp (starvation lipoprotein) and hdeD (acid-resistance membrane protein) (34, 50, 64, 89). The down-regulated transcriptional unit gadE-mthF also related to low pH, including gene gadE (GadE transcriptional activator), mth (subunit of MthF-Tolc multidrug efflux transport system) and mth 106 (subunit of MthF-Tolc multidrug efflux transport system) (64). Quite a few of down-regulated stress related genes involved in response to pH. Even although ZnO NPs can cause genotoxicity, some genes related to DNA repair and detoxification were still down-regulated, such as gene rqu (Holliday Junction nuclease). The repressed gene uer, coding DNA-dependent ATPase I and helicase II, can also protect DNA from degradation. Gene ksgA was downregulated, coding S-adenosylmethionine-6-N',N'-adenosyl (rRNA) dimethyltransferase, and its expression is positively regulated by growth rate (74). Meanwhile, some genes functioning in nucleotide metabolism were also down-regulated. Gene rutC (endoribonuclease), rutD (hydrolase) and rutE (nitroreductase) in the operon rutABCDEF G were repressed. These genes function in the pathway for pyrimidine degradation pathway, and the operon is found under the control of nitrogen regulatory protein (107). The transcriptional unit xdhABC was also down-regulated, including xdhA, xth, and xdhC, which consist of xanthine dehydrogenase, and xanthine dehydrogenase plays a role in purine salvage (101). Some genes involved in detoxification were repressed by ZnO NPs. For example, repressed gene yiiM codes a protein involved in base analog detoxification (54). Moreover, in the transcriptional unit yan-dinJ, coding the Yan-DinJ toxin-antitoxin system, Yan functions in reducing protein synthesis and inhibiting growth, when DinJ alleviates that phenotype, acting as the antitoxin (67). Both of these two genes were down-regulated in the treatment of ZnO NPs. A few genes about cell division, cell motion 107 and biofilm formation were also repressed. For example, repressed gene yihA, coding cell division protein, is essential for normal cell division (6, 23). RNA polemrase sigma F factor coded by gene fliA was also down-regulated, and it is related to flagellin production, motility and biofilm formation (10, 60). Gene expression profiles following Si02 treatment. In SiOz treatment, 5924 genes ranged in the expression fold between 0.40 and 2.05, which was much narrower than those in these other two treatments (Fig 5.1). Based on the available information in the databases of Clusters of Orthologous Groups of proteins (COGS), fifty-two down-regulated genes in the response to SiOz NPs were grouped into 17 functional categories, among which “transcription” (11.5%) and “Not in COGS” (30.8%) were the two of the biggest dominant categories. Other large categories included “Inorganic ion transport and metabolism” (7.7%), “coenzyme transport and metabolism” (5.8%), “Cell wall/membrane/envelope biogenesis” (5.8%), “General function” (5.8%), “Function unknown” (5.8%), “defense mechanisms” (5.8%). However, only 2 genes were down-regulated in the categories “transcription” and “function unknown”. “Transportation” is always one of the biggest categories in all the treatments, and for Si02 NP exposure 10 out of 52 down—regulated genes were in this category, covering transporters for carbohydrate, amino acid and metal ions, such as gene ych, yeeO, yaaU, yiaV, and ych (14). Gene ych, functioning as ferrous iron transporter, was repressed probably due to osmotic up-shift that may be caused by membrane damage from ROS, or 108 anaerobic conditions (48, 95). Meanwhile, gene napC, involving in transportation and anaerobic respiration, was also repressed. Gene cyoB is another repressed gene involved in respiration, and it codes cytochrome terminal oxidase subunit 1, a terminal oxidase in the respiratory chain used under high oxygen growth conditions as a proton pump (77). Moreover, both genes phnE and ccmB are ABC transporters, functioning as the integral membrane protein and the membrane protein of type 1 cytochrome c biogenesis system respectively (77, 94). “Transcription” is another big category in the treatment of Si02 NPs. The down-regulated gene ngJ was considered as a positive DNA-binding transcriptional regulator of transport and utilization of the aromatic B-glucosides arbutin and salicin. YgeK was repressed as a predicted response regulators of two-component regulatory systems (62, 103). Gene ascG was also down-regulated, which can repress the expression of ascF B involving in transport and utilization of the glucosidse suargars arbutin, salicin, and cellibiose (39). Exposure to NPs can cause some cellular stress and damage, such as oxidative stress and genotoxicity by ROS, and they also influence the genes related to stress condition and DNA repair. For instance, gene yodA, which can bind divalent metal ions including cadmium, zinc, and nickel and is proposed to be a generalized stress factor, was repressed (24). Meanwhile, gene yedQ was also found to be down-regulated, which responds to a number of stress conditions and involve in cellulose biosynthesis (96). 109 Gene xisE was repressed, functioning as DNA recombination, when gene cca was also down-regulated, involving in RNA repair (56). There were a few of transcriptional units down-regulated dLuing the exposure to Si02 NPs. The transcriptional unit nhaAR consists of gene nhaA, a sodium ion/proton antiporter that uses the proton electrochemical gradient to expel sodium ions from the cytoplasm and functions primarily in the adaptation to high salinity at alkaline pH, and nhaR, a transcriptional activator. Moreover, nhaA is also believed to be responsible for adaptation to alkaline pH when sodium is available (49). In the transcriptional unit hcaEFCBD, hcaE coding a subunit of 3-phenylpropionate dioxygenase, hcaF coding [3 subunit of 3-phenylpropionate dioxygenase, and hcaB coding 3-phenylpropionate-2',3'-dihydrodiol dehydrogenase were repressed, functioning in utilize aromatic acids as carbon and energy sources (15). Gene expression profile following Ti02 treatment. Totally 5940 genes ranged in the expression fold from 0.22 to 7.25. In Figure 5.1, globally 11 down-regulated genes in the TiOz treatment were grouped into 7 functional categories, including “energy production and conversion” (27.3%), “coenzyme transport and metabolism” (18.2%), “general function” (18.2%), “cell motility” (9.1%), “inorganic ion transport and metabolism” (9.1%), “signal transduction mechanisms” (9.1%) and “not in COGS” (9.1%). Similarly, twenty-eight up-regulated genes were grouped into 10 functional categories. Besides “amino acid transport and metabolism” (32.1%) and “not in COGS” 110 (17.9%), most of induced genes belonged to the categories of “carbohydrate transport and metabolism” (10.7%), “inorganic ion transport and metabolism” (10.7%), “energy production and conversion” (7.1%), “transcription” (7.1%). Most of repressed genes in the treatment of TiOz were related to transportation and cofactor synthesis. Gene yejF was down-regulated, coding an ATP-binding protein of a ABC transporter that transports peptides through membrane, and its expression was reported to be regulated by small RNA at the mRNA level (5, 71, 82). The repressed genes moaA and moaB were in the transcriptional unit moaABCDE, functioning in Mo-molybdopterin cofactor biosynthetic process (4). The gene deB, coding dimethyl sulfoxide (DMSO) reductase chain B, was also down-regulated. Moreover, one of DMSO reductase paralogues was downregulated, including gene ynfE (oxidoreductase subunit paralog of DmsA), yn/F (oxidoreductase subunit), ynfG (oxidoreductase, Fe-S subunit paralog of DmsB), and yan (DMSO reductase anchor subunit paralog of DmsC) in the same operon, which utilize DMSO as a terminal oxidant and function in energy production and electron transferring (61). Both the DMSO reductase and its paralogue catalyze in anaerobic electron transport chain, and they contain iron ions and molybdenum cofactor. Their down-regulation probably relate to the repression of molybdopterin biosynthesis. Among up-regulated genes, transportation (32.1%) and amino acid biosynthesis and catabolism (42.9%) were the biggest categories. For the instance, gene ybaE was 111 up-regulated, functioning as ‘ABC transporter. In the transcriptional unit of glucitol/sorbitol phosphotransferase system (PTS) permease, genes erA (glucitol/sorbitol-specific enzyme IIC component of PTS), srlB (glueitol/sorbitol-specific enzyme IIA component of PTS) and erE (glucitol/sorbitol-Specific enzyme IIB component of PTS) were up-regulated, which involve in carbohydrate transport via glycolysis (76). It was reported that GutM and GutR are positive and negative transcriptional regulators of srl Operon expression respectively, and the operon expression is also under the control of the cyclic AMP-cyclic AMP receptor protein (CRP) complex (102). The transcriptional unit of modABC is another induced operon, coding ModA, ModB and ModC as subunits of molybdate ABC transporter, and their expression was found to be regulated by a molybdate-dependent repressor protein (36, 37, 100). For amino acid biosynthesis-and catabolism, the transcriptional unit gltBDF were induced, coding a glutamate synthase that consists of the large (GltB) and small (GltD) subunits, and a protein that is implied in the self-operon regulation (GltF). It was reported that all of these three subunits were induced by exposure to Ti02 NPS, and they involve in glutamate biosynthesis (18, 19). The activation of this operon could be due to Lrp (global modulator in response to. leucine levels), IHF (global DNA bending proteins) and GadE (regulator for the resistance to low pH) (43, 98). The thrLABC operon codes four enzymes involved in threonine and homoserine biosynthesis, all of which were up-regulated. Moreover, the tchBCDEFG operon was induced, which codes L-serine deaminase III (Tch), endoribonuclease (Tch), 2-ketobutyrate formate-lyase/pyruvate 112 formate-lyase (Tch), propionate kinase (Tch), threonine STP transporter (Tch), catabolic threonine dehydratase (Tch), and Tch transcriptional activator (Tch), involving amino acid catabolism-transport, and anaerobic respiration (27, 41, 72, 86, 90). This operon can be induced under anaerobiosis, but it is repressed by glucose (83). The genes 1'le and nmpC were also up-regulated, relating to amino acid biosynthesis and ion transporter activity respectively. Gene ach, which functions in two-component signal transduction system and aerobic respiration, was induced. It involves in amino acid catabolism, transportation, and anaerobic respiration (27, 41, 90). Validation of microarray data by quantitative PCR. In Table 5.1, quantitative PCR for 10 selected genes was used to test the robustness of the microarray hybridization data, which included two highly repressed genes (yheN and hdeB), two moderately repressed genes (hyaB and manX), a slightly induced gene (mltC), three slightly induced genes (atpD, sdhC and mopB), one moderately induced gene (ipr) and another moderately induced gene with high variations (glpA). Gene ch was used as a reference gene for comparative CT method; A high correlation (R2=0.95) was obtained between microarray analysis and quantitative PCR analysis. Discussion Aerobic/anaerobic transition in the NP solution. Among the influenced genes, there were quite a few of them related to aerobic/anaerobic conditions, especially in ZnO treatment. In ZnO NP treatment, most of repressed genes about aerobic/anaerobic 113 conditions were related to anaerobic respiration, while there were both of genes about aerobic and anaerobic respiration induced. For example, in the transcriptional unit appCBA-yccB, genes appC (subunit I of cytochrome bd—II terminal oxidase), appB (subunit II of cytochrome bd-II terminal oxidase) and appA (acid phosphatase) was down-regulated, and this operon codes cytochrome bd-II terminal oxidase, which was not expressed under normal conditions of growth. Meanwhile, gene cyoA, cyoB and cyoD in transcriptional unit cyoABCDE were induced, which code cytochrome b0 terminal oxidase used under high oxygen growth conditions. Another example is both aerobic and anaerobic glycerol-3-phosphate dehydrogenases, functioning in glycerol metabolism and amino acid transport, as well as providing the proton-motive force for ATP synthesis (55). The anaerobic glycerol-3-phosphate dehydrogenase consists of GlpA (large subunit), GlpB (membrane anchor subunit) and GlpC (small subunit), and aerobic one is coded by gene glpD, both of which was induced by ZnO NPS. Moreover, gene tch coding catabolic threonine dehydratase was induced by ZnO NPS, and it was reported to be only expressed during anaerobic growth in the absence of glucose, suggesting the growth during a transition from aerobic to anaerobic condition. Meanwhile, gene doc was repressed, coding heme-regulated phosphodiesterase and acting as a direct sensor of oxygen (25, 66). Gene fucO, coding L-1,2-propanediol oxidoreductase, induced by growth on L-fucose under both aerobic and anaerobic conditions, but inactivation of the enzyme under aerobic conditions (16). In TiOz NP suspensions, anaerobic DMSO reductaSe and its paralogue were down-regulated while the operon tchBCDEF G was 114 induced, which involves in L-serine degradation and anaerobic amino acid metabolism. The repression of some anaerobic pathways and the induction of some anaerobic/aerobic pathways suggest that aerobic/anaerobic transition may occur in NP suspension. It was also reported that fullerene can inhibit E. coli respiration and dioxygen consumption (63). The function of iron during the exposure to NPS. Iron ions play an important role for microorganisms. During the exposure to NPS, they can serve as a reducing agent and repair oxidative damage. For example, gene dps, coding Dps complex that can sequester iron, was repressed by ZnO NPS, although it can bind to DNA and protect them from oxidative damage (106). It was reported that protein Dps is very abundant in stationary-phase and in normal starvation response, which is contrary to our conditions (45). In the category “transportation”, most of repressed genes related to carbohydrate and amino acid transport, while 25% of induced transporter genes function in ion transport. In fact, ROS can attack sulfur-iron cluster , and lead to Fenton chemistry to generate more ROS (46). Many oxidoreductases, dehydrogenases, hydrogenases, oxidases, and reductases contain iron-sulfur clusters or bind iron ions, and that is why their expression was regulated by NPS. In the operon squBCDSE, gene sufB (component of SufBCD cysteine desulfurase activator complex), squ (component of SufBCD complex), sujD (component of SufBCD cysteine desulfurase activator complex), sufS (subunit of selenocysteine lyase), and squ (sulfur acceptor) were repressed. This operon synthesizes the enzymes in a secondary pathway of iron-sulfur cluster assembly, which is related to intracellular availability’of iron ions (88). This may be related to oxidation of 115 sulfur atoms by ROS (46). The bacterial tress-relatedgenes during NP exposures. The stress-related genes were found in both of down-regulated and up-regulated gene profiles. Some transcription regulators can sense the stress and influence the expression of other genes. There were more induced genes about response to stress than repressed ones, such as those coding starvation proteins, acid resistant proteins and heat shock proteins. Quite a few genes responding to osmotic stress were up-regulated, maybe due to ROS induction. ROS can oxidize fatty acids of phospholipids on membrane, and cause lipid peroxidation. This process can change permeability and fluidity of bacterial membrane, making bacteria more vulnerable to osmotic stress and interrupt nutrient uptake (17). Conclusion The influence on the gene expression of E. coli increased generally, from Si02 to TiOz to ZnO NPS. The number of induced and repressed genes was the highest in ZnO NPS, and for all the three treatment the number of repressed genes is higher than the repressed ones. ROS induction was evidenced by microarray data, such as the over-expression of osmotic stress genes that may be due to the alteration of permeability and fluidity caused by lipid peroxidation, and expression change of iron ion transporter led by attacking iron-sulfur cluster. Both of them result in the damage from ROS, which also hinder nutrient uptake and induce DNA pair genes. It was also demonstrated that ZnO NPS influenced the microbial respiration. The genes related to anaerobic and aerobic 116 respiration was induced, while only genes about anaerobic respiration were repressed. The microarray data demonstrated that NPS regulated the genes in various categories, and adjusted the pathways and the metabolisms of E. coli. 117 3N omN toamcab .8328.» can £030.53 .wEu—oEab 83.—comb:— ”D mEflgnooE cozoscmcah .a:w_m fl. 9505?: cococzm “m bco :ozflvoa cases—d 35:00 um .m:onfiao can team—Eb .m_m05:>mo_n mug—09805 Eucooom “O Em=onSoE van ton—v.53 :2 0:899: um museums-U do>oEB 530.5 .cozmocmvoE 3:253:58.”— ”0 3:22: :00 Hz mica—owe}. oao_o>:o\o:aBEoE\=a3 :00 ”E bun—o.— Uca GOES—55000. .comumflaod ”A DIIIIIEDIIII cozatomcmfi. ”v— flmocowoB van 0.583% .mEOmont doze—map; ”a Sam—00505 van roam—Sb Ea: ”— Emzonfioe use toga—5.: thucooU ”I EmzonSoE v5 teamcab 035.309ch ”0 CHE—0830:.— Uca ton—mach ounce—o: Z ”L Efl-OQSOE can ton—mafia Boa OEE< ”m Bflllllll choEtea oEOmoEoEo 62320 :00 ._ob:oo Bozo :00 ”Q r: o F: i- 0 > r: o 0 E r: 2 OJ 0 3 '0 o h a. E 0 z: m U 00—. omw OON omN iaqrunu was 119 References 1. Abaibou, H., G. Giordano, and M. A. MandrandBerthelot. 1997. Suppression of Escherichia coli formate hydrogenlyase activity by trimethylamine N-oxide is due to drainage of the inducer formate. Microbiology-Uk 143:2657-2664. 2. Abdel-Hamid, A. M., M. M. Attwood, and J. R. Guest. 2001. Pyruvate oxidase contributes to the aerobic growth efficiency of Escherichia coli. Microbiology-ng 147:1483-1498. 3. Adams, L. K., D. Y. Lyon, and P. J. J. Alvarez. 2006. Comparative eco-toxicity of nanoscale Ti02, Si02, and ZnO water suspensions. Water Research 40:3527-3532. 4. Anderson, L. A., E. McNairn, T. Leubke, R. N. Pan, and D. H. Boxer. 2000. ModB-dependent molybdate regulation of the molybdenum cofactor operon moa in Escherichia coli. Journal of Bacteriology 182:7035-7043. 5. Antal, M., V. Bordeau, V. Douchin, and B. F elden. 2005. A small bacterial RNA regulates a putative ABC transporter. Journal of Biological Chemistry 280:7901-7908. 6. Arigoni, F., F. Talabot, M. Peitsch, M. D. Edgerton, E. Meldrum, E. Allet, R. Fish, T. Jamotte, M. L. Curchod, and H. Loferer. 1998. A genome-based approach for the identification of essential bacterial genes. Nature Biotechnology 16:851-856. 7. Aruoja, V., H. C. Dubourguier, K. Kasemets, and A. Kahru. 2008. Toxicity of nanoparticles of CuO, ZnO and Ti02 to microalgae pesudokirchneriella subcapitata. Science of he Total Environment 407: 1461-1468. 8. Ashwell, G, A. J. Wahba, and J. Hickman. 1960. Uronic acid metabolism in bacterial. Purification and properties of uronic acid isomerase in Escherichia-coli. Journal of Biological Chemistry 235: 1559-1565. 9. Azizan, A., and P. N. Black. 1994. Use of Transposon Tnphoa to Identify Genes 120 10. ll. 12. 13. 14. 15. 16. 17. for Cell-Envelope Proteins of Escherichia-Coli Required for Long-Chain Fatty-Acid Transport - the Periplasmic Protein Tsp Potentiates Long-Chain Fatty-Acid Transport. Journal of Bacteriology 176:6653-6662. Barrios, A. F. G., R. J. Zuo, D. C. Ren, and T. K. Wood. 2006. Hha, YbaJ, and OmpA regulate Escherichia coli K12 biofilm formation and conjugation plasmids abolish motility. Biotechnology and Bioengineering 93: 1 88-200. Bertero, M. G., R. A. Rothery, M. Palak, C. Ron, D. Lim, F. Blasco, J. H. Weiner, and N. C. J. Strynadka. 2003. Insights into the respiratory electron transfer pathway from the structure of nitrate reductase A. Nature Structural Biology 10:681-687. Brondijk, T. H. C., D. Fiegen, D. J. Richardson, and J. A. Cole. 2002. Roles of NapF, NapG and NapH, subunits of the Escherichia coli periplasmic nitrate reductase, in ubiquinol oxidation. Molecular Microbiology 44:245-255. Brondijk, T. H. C., A. Nilavongse, N. Filenko, D. J. Richardson, and J. A. Cole. 2004. NapGH components of the periplasmic nitrate reductase of Escherichia coli K-12: location, topology and physiological roles in quinol oxidation and redox balancing. Biochemical Journal 379:47-55. Brown, M. H., I. T. Paulsen, and R. A. Skurray. 1999. The multidrug efflux protein NorM is a prototype of a new family of transporters. Molecular Microbiology 31 :394-395. Burlingame, R. P., L. Wyman, and P. J. Chapman. 1986. Isolation and characterization of Escherichia-coli mutants defective for phenylpropionate degradation. Journal of Bacteriology 168:55-64. Cabiscol, E., E. Hidalgo, J. Badia, L. Baldoma, J. Q. Ros, and J. Aguilar. 1990. Oxygen regulation of L-1,2-propanediol oxidoreductase activity in Escherichia-coli. Journal of Bacteriology 172:5514-5515. Cabiscol, E., J. Tamarit, and J. Ros. 2000. Oxidative stress in bacteria and protein damage by reactive oxygen Species. International Microbiology 3:3-8. 121 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. Castano, 1., F. Bastarrachea, and A. A. Covarrubias. 1988. Gltbdf operon of Escherichia-coli. Journal of Bacteriology 170:821-827. Castano, 1., N. Flores, F. Valle, A. A. Covarrubias, and F. Bolivar. 1992. Gltf, a member of the gltbdf operon of Escherichia-coli, is involved in nitrogen-regulated gene-expression. Molecular Microbiology 6:2733-2741. Chang, Y. Y., A. Y. Wang, and J. E. Cronan. 1994. Expression of Escherichia-coli pyruvate oxidase (Poxb) depends on the sigma-factor encoded by the rpos(katf) gene. Molecular Microbiology 11:1019-1028. Conter, A., J. P. Bouche, and M. Dassain. 1996. Identification of a new inhibitor of essential division gene ftsZ as the kil gene of defective prophage Rac. Journal of Bacteriology 178:5100-5104. CourcelIe, J., and P. C. Hanawalt. 2003. RecA-dependent recovery of arrested DNA replication forks. Annual Review of Genetics 37:611-646. Dassain, M., A. Leroy, L. Colosetti, S. Carole, and J. P. Bouche. 1999. A new essential gene of the 'minimal genome‘ affecting cell division. Biochimie 81:889-895. David, G., K. Blondeau, M. Schiltz, S. PeneI, and A. Lewit-Bentley. 2003. YodA from Escherichia coli is a metal-binding, lipocalin-like protein. Journal of Biological Chemistry 278:43728-43735. Delgado-Nixon, V. M., G. Gonzalez, and M. A. Gilles-Gonzalez. 2000. Dos, a heme-binding PAS protein from Escherichia coli, is a direct oxygen sensor. Biochemistry 39:2685-2691. Dufour, A., R. B. F urness, and C. Hughes. 1998. Novel genes that upregulate the Proteus mirabilis flhDC master operon controlling flagellar biogenesis and swarming. Molecular Microbiology 29:741-751. Egan, R. M., and A. T. Phillips. 1977. Requirements for induction of biodegradative threonine dehydratase in Escherichia-coli. Journal of Bacteriology 122 28. 29. 30. 31. 32. 33. 34. 35. 36. 132:370-376. Egler, M., C. Grosse, G. Grass, and D. H. Nies. 2005. Role of the extracytoplasmic function protein family sigma factor RpoE in metal resistance of Escherichia coli. Journal of Bacteriology 187 :2297-2307. Engineering, T. R. S. a. t. R. A. o. 2004. Nanoscience and nanotechnologies: opportunities and uncertainties. Eppler, T., P. Postma, A. Schutz, U. Volker, and W. Boos. 2002. Glycerol-3-phosphate-induced catabolite repression in Escherichia coli. Journal of Bacteriology 184:3044-3052. F eissner, R. E., C. L. Richard-Fogal, E. R. Frawley, and R. G. Kranz. 2006. ABC transporter-mediated release of a haem chaperone allows cytochrome c biogenesis. Molecular Microbiology 61:219-231. Franklin, N. M., N. J. Rogers, S. C. Apte, G E. Batley, G. E. Gadd, and P. S. Casey. 2007. Comparative toxicity of nanoparticulate ZnO, bulk ZnO, and ZnC12 to a freshwater microalga (Pseudokirchneriella subcapitata): The importance of particle solubility. Environmental Science & Technology 41:8484-8490. lFu, G. F., P. S. Vary, and C. T. Lin. 2005. Anatase Ti02 nanocomposites for antimicrobial coatings. Journal of Physical Chemistry B 109:8889-8898. Gajiwala, K. S., and S. K. Burley. 2000. HDEA, a periplasmic protein that supports acid resistance in pathogenic enteric bacteria. Journal of Molecular Biology 295:605-612. Garland, P. B., J. A. Downie, and B. A. Haddock. 1975. Proton translocation and respiratory nitrate reductase of Escherichia-coli. Biochemical Journal 152:547-559. Grunden, A. M., R. M. Ray, J. K. Rosentel, F. G. Healy, and K. T. Shanmugam. 1996. Repression of the Escherichia coli modABCD (molybdate transport) operon by ModE. Journal of Bacteriology 178:735-744. 123 37. 38. 39. 40. 41. 42. 43. 44. 45. Grunden, A. M., W. T. Self, M. Villain, J. E. BIanck, and K. T. Shanmugam. 1999. An analysis of the binding of repressor protein ModE to modABCD (Molybdate transport) operator/promoter DNA of Escherichia coli. Journal of Biological Chemistry 274:24308-24315. Gutheil, W. G., E. Kasimoglu, and P. C. Nicholson. 1997. Induction of glutathione-dependent formaldehyde dehydrogenase activity in Escherichia coli and Hemophilus influenza. Biochemical and Biophysical Research Communications 238:693-696. Hall, B. G., and L. Xu. 1992. Nucleotide-sequence, function, activation, and evolution of the cryptic asc operon of Escherichia-coli K12. Molecular Biology and Evolution 9:688-706. Herring, C. D., and F. R. Blattner. 2004. Global transcriptional effects of a suppressor tRNA and the inactivation of the regulator frmR. Journal of Bacteriology 186:6714-6720. Hesslinger, C., S. A. Fairhurst, and G. Sawers. 1998. Novel keto acid formate-lyase and propionate kinase enzymes are components of an anaerobic pathway in Escherichia coli that degrades L-threonine to propionate. Molecular Microbiology 27 :477-492. Hirakawa, H., K. Nishino, T. Hirata, and A. Yamaguchi. 2003. Comprehensive studies of drug resistance mediated by overexpression of response regulators of two-component Signal transduction systems in Escherichia coli. Journal of Bacteriology 185:1851-1856. Hommais, F., E. Krin, J. Y. Coppee, C. Lacroix, E. Yeramian, A. Danchin, and P. Bertin. 2004. GadE (YhiE): a novel activator involved in the response to acid environment in Escherichia coli. Microbiology-ng 150:61-72. Horlacber, R., and W. Boos. 1997. Characterization of TreR, the major regulator of the Escherichia coli trehalose system. Journal of Biological Chemistry 272:13026-13032. Ilari, A., P. Ceci, D. Ferrari, G. L. Rossi, and E. Chiancone. 2002. Iron 124 46. 47. 48. 49. 50. 51. 52. 53. incorporation into Escherichia coli Dps gives rise to a ferritin-like microcrystalline core. Journal of Biological Chemistry 277 :37619-37623. Imlay, J. A. 2003. Pathways of oxidative damage. Annual Review of Microbiology 57:395-418. Iuchi, S., A. Aristarkhov, J. M. Dong, J. S. Taylor, and E. C. C. Lin. 1994. Effects of nitrate respiration on expression of the Arc-controlled operons encoding succinate-dehydrogenase and flavin—linked L-lactate dehydrogenase. Journal of Bacteriology 176: 1695-1 701. Kang, Y. S., K. D. Weber, Q. Yu, P. J. Kiley, and F. R. Blattner. 2005. Genome-wide expression analysis indicates that FNR of Escherichia coli K-12 regulates a large number of genes of unknown function. Journal of Bacteriology 187:1135-1160. Kedrov, A., and D. J. Muller. 2006. Characterizing folding, structure, molecular interactions and ligand gated activation of single sodium/proton antiporters. Naunyn-Schmiedebergs Archives of Pharmacology 372:400-412. Kern, R., A. Malki, J. Abdallah, J. Tagourti, and G. Richarme. 2007. Escherichia coli HdeB is an acid stress chaperone. Journal of Bacteriology 189:603-610. King, P. W., and A. E. Przybyla. 1999. Response of hya expression to external pH in Escherichia coli. Journal of Bacteriology 181:5250-5256. Klaine, S. J., P. J. J. Alvarez, G. E. Batley, T. F. Fernandes, R. D. Handy, D. Y. Lyon, S. Mahendra, M. J. McLaughlin, and J. R. Lead. 2008. Nanomaterials in the environment: Behavior, fate, bioavailability, and effects. Environmental Toxicology and Chemistry 27:1825-1851. Kolesnikow, T., I. Schroder, and R. P. Gunsalus. 1992. Regulation of Nark gene-expression in Escherichia-coli in response to anaerobiosis, nitrate, iron, and molybdenum. Journal of Bacteriology 174:7104-7111. 125 54. 55. 56. 57. 58. 59. 60. 61. 62. Kozmin, S. G., P. Leroy, Y. I. Pavlov, and R. M. Schaaper. 2008. chX and yiiM, two novel determinants for resistance of Escherichia coli to N-hydroxylated base analogues. Molecular Microbiology 68:51-65. Kuritzkes, D. R., X. Y. Zhang, and E. C. C. Lin. 1984. Use of phi-(glp-lac) in studies of respiratory regulation of the Escherichia-coli anaerobic sn-glycerol-3-phosphate dehydrogenase genes (glpab). Journal of Bacteriology 157:591-598. Kuznetsova, E., M. Proudfoot, S. A. Sanders, J. Reinking, A. Savchenko, C. H. Arrowsmith, A. M. Edwards, and A. F. Yakunin. 2005. Enzyme genomics: Application of general enzymatic screens to discover new enzymes. Ferns Microbiology Reviews 29:263-279. Laurinavichene, T. V., N., A. Zorin, and A. A. Tsygankov. 2002. Effect of redox potential on activity of hydrogenase 1 and hydrogenase 2 in Escherichia coli. Archives of Microbiology 178:437-442. Lee, L. J., J. A. Barrett, and R. K. Poole. 2005. Genome-wide transcriptional response of chemostat-cultured Escherichia coli to zinc. Journal of Bacteriology 187:1124-1134. Limon, A., E. Hidalgo, and J. Aguilar. 1997. The aldA gene of Escherichia coli is under the control of at least three transcriptional regulators. Microbiology-Uk 143:2085-2095. Liu, X. Y., and P. Matsumura. 1995. An alternative sigma-factor controls transcription of flagellar class-III operons in Escherichia-coli - Gene sequence, overproduction, purification and characterization. Gene 164:81-84. Lubitz, S. P., and J. H. Weiner. 2003. The Escherichia coli ynfEFGHI operon encodes polypeptides which are paralogues of dimethyl sulfoxide reductase (DmSABC). Archives of Biochemistry and Biophysics 418:205-216. Madhusudan, S., A. Paukner, Y. Klingen, and K. Schnetz. 2005. Independent regulation of H-NS-mediated silencing of the bgl operon at two levels: upstream by BgIJ and LeuO and downstream by DnaKJ. Microbiology-ng 126 63. 64. 65. 66. 67. 68. 69. 70. 151:3349-3359. Mashino, T., K. Okuda, T. Hirota, M. Hirobe, T. Nagano, and M. Mochizuki. 1999. Inhibition of E-coli growth by fullerene derivatives and inhibition mechanism. Bioorganic & Medicinal Chemistry Letters 9:2959-2962. Masuda, N., and G. M. Church. 2003. Regulatory network of acid resistance genes in Escherichia coli. Molecular Microbiology 48:699-712. Maurer, L. M., E. Yohannes, S. S. Bondurant, M. Radmacher, and J. L. Slonczewski. 2005. pH regulates genes for flagellar motility, catabolism, and oxidative stress in Escherichia coli K-12. Journal of Bacteriology 187:304-319. Mejean, V., C. Iobbinivol, M. Lepelletier, G. Giordano, M. Chippaux, and M. C. Pascal. 1994. Tmao anaerobic respiration in Escherichia-coli - Involvement of the Tor operon. Molecular Microbiology 11:1169-1179. Motiejunaite, R., J. Armalyte, A. Markuckas, and E. Suziedeliene. 2007. Escherichia coli dinJ-yan genes act as a toxin-antitoxin module. Ferns Microbiology Letters 268: 1 12-1 19. ' Muela, A., C. Seco, E. Camafeita, I. Arana, M. Orruno, J. A. Lopez, and I. Barcina. 2008. Changes in Escherichia coli outer membrane subproteome under environmental conditions inducing the viable but nonculturable state. Ferns Microbiology Ecology 64:28-36. Nakamura, K., M. Yamaki, M. Sarada, S. Nakayama, C. R. T. Vibat, R. B. Gennis, T. Nakayashiki, H. Inokuchi, S. Kojima, and K. Kita. 1996. Two hydrophobic subunits are essential for the heme b ligation and functional assembly of complex 11 (succinate-ubiquinone oxidoreductase) from Escherichia coli. Journal of Biological Chemistry 271:521-527. Neal, A. L. 2008. What can be inferred from bacterium-nanoparticle interactions about the potential consequences of environmental exposure to nanoparticles? Ecotoxicology 17:362-371. 127 71. 72. 73. 74. 75. 76. 77. 78. 79. Novikova, M., A. Metlitskaya, K. Datsenko, T. Kazakov, A. Kazakov, B. Wanner, and K. Severinov. 2007. The Escherichia coli yej transporter is required for the uptake of translation inhibitor microcin C. Journal of Bacteriology 189:8361-8365. Ogawa, W., T. Kayahara, M. Tsuda, T. Mizushima, and T. Tsuchiya. 1997. Isolation and characterization of an Escherichia coli mutant lacking the major serine transporter, and cloning of a serine transporter gene. Journal of Biochemistry 122: 1241-1245. Park, S. J., C. P. Tseng, and R. P. Gunsalus. 1995. Regulation of succinate-dehydrogenase (Sdhcdab) operon expression in Escherichia-coli in response to carbon supply and anaerobiosis - Role of Area and Fnr. Molecular Microbiology 15:473-482. Pease, A. J., B. R. Roa, W. Luo, and M. E. Winkler. 2002. Positive growth rate-dependent regulation of the pdxA, ksgA, and pde genes of Escherichia coli K-12. Journal of Bacteriology 184:1359-1369. PeIIicer, M. T., A. S. Lynch, P. De Wulf, D. Boyd, J.’ Aguilar, and E. C. C. Lin. 1999. A mutational study of the ArcA-P binding sequences in thelaldA promoter of Escherichia coli. Molecular and General Genetics 261:170-176. Postma, P. W., J. W. Lengeler, and G. R. Jacobson. 1993. Phosphoenolpyruvate - carbohydrate phosphotransferase systems of bacteria. Microbiological Reviews 57:543-594. Puustinen, A., M. Finel, T. Haltia, R. B. Gennis, and M. Wikstrtim. 1991. Properties of the two terminal oxidases of Escherichia coli. Biochemistry 30:3936-3942. Richard, D. J., G. Sawers, F. Sargent, L. McWalter, and D. H. Boxer. 1999. Transcriptional regulation in response to oxygen and nitrate of the operons encoding the [NiFe] hydrogenases l and 2 of Escherichia coli. Microbiology-Uk 145:2903-2912. Richey, D. P., and E. C. C. Lin. 1972. Importance of Facilitated Diffusion for 128 80. 81. 82. 83. 84. 85. 86. 87. Effective Utilization of Glycerol by Escherichia-Coli. Journal of Bacteriology 112:784-&. Rincon, A. G., and C. Pulgarin. 2004. Bactericidal action of illuminated Ti02 on pure Escherichia coli and natural bacterial consortia: post-irradiation events in the dark and assessment of the effective disinfection time. Applied Catalysis B-Environmental 49:99-1 12. Rincon, A. G., and C. Pulgarin. 2004. Effect of pH, inorganic ions, organic matter and H202 on E-coli K12 photocatalytic inactivation by Ti02 - Implications in solar water disinfection. Applied Catalysis B-Environmental 51:283-302. Saurin, W., M. Hofnung, and E. Dassa. 1999. Getting in or out: Early segregation between importers and exporters in the evolution of ATP-binding cassette (ABC) transporters. Journal of Molecular Evolution 48:22-41. Sawers, G. 2001. A novel mechanism controls anaerobic and catabolite regulation of the Escherichia coli tdc operon. Molecular Microbiology 39: 1285-1298. Seputiene, V., D. Motiejunas, K. Suziedelis, H. Tomenius, S. Normark, O. Melefors, and E. Suziedeliene. 2003. Molecular characterization of the acid-inducible asr gene of Escherichia coli and its role in acid stress response. Journal of Bacteriology 185:2475-2484. Sharma, R., C. Reusing, B. P. Rosen, and B. Mitra. 2000. The ATP hydrolytic activity of purified ZntA, a Pb(II)/Cd(II)/Zn(II)-translocating ATPase from Escherichia coli. Journal of Biological Chemistry 275:3873-3878. Sumantran, V. N., H. P. Schweizer, and P. Datta. 1990. A novel membrane-associated threonine permease encoded by the tch gene of Escherichia-coli. Journal of Bacteriology 172:4288-4294. T Hayes, E., J. C. Wilks, P. Sanfilippo, E. Yohannes, D. P. Tate, B. D Jones, M. D Radmacher, S. S. BonDurant, and J. L. Slonczewski. 2006. Oxygen limitation modulates pH regulation of catabolism and hydrogenases, multidrug transporters, and envelope composition in Escherichia coli K-12. Bmc 129 88. 89. 90. 91. 92. 93. 94. 95. 96. Microbiology 6:-. Takahashi, Y., and U. Tokumoto. 2002. A third bacterial system for the assembly of iron-sulfur clusters with homologs in archaea and plastids. Journal of Biological Chemistry 277:28380-28383. Tucker, D. L., N. Tucker, and T. Conway. 2002. Gene expression profiling of the pH response in Escherichia coli. Journal of Bacteriology 184:6551-6558. Umbarger, H. E., and B. Brown. 1957. Threonine deamination in Escherichia-coli.2. Evidence for 2 L-threonine deaminases. Journal of Bacteriology 73:105-112. Via, P., J. Badia, L. Baldoma, N. Obradors, and J. Aguilar. 1996. Transcriptional regulation of the Escherichia coli rhaT gene. Microbiology-Uk 142:1833-1840. Wang, H. N., and R. P. Gunsalus. 2003. Coordinate regulation of the Escherichia coli formate dehydrogenase fdnGHI and fth genes in response to nitrate, nitrite, and formate: Roles for NarL and NarP. Journal of Bacteriology 185:5076-5085. Wang, H. N., C. P. Tseng, and R. P. Gunsalus. 1999. The napF and narG nitrate reductase operons in Escherichia coli are differentially expressed in response to submicromolar concentrations of nitrate but not nitrite. Journal of Bacteriology 181:5303-5308. Wanner, B. L., and W. W. Metcalf. 1992. Molecular genetic-studies of a 10.9-Kb operon in Escherichia-coli for phosphonate uptake and biodegradation. F ems Microbiology Letters 100: 133-139. Weber, A., and K. Jung. 2002. Profiling early osmostress-dependent gene expression in Escherichia coli using DNA microarrays. Journal of Bacteriology- 184:5502-5507. Weber, H., C. Pesavento, A. Possling, G. Tischendorf, and R. Hengge. 2006. 130 97. 98. 99. 100. 101. 102. 103. 104. Cyclic-di-GMP-mediated signalling within the sigma(S) network of Escherichia coli. Molecular Microbiology 62: 1014-1034. Weissenborn, D. L., N. Wittekindt, and T. J. Larson. 1992. Structure and regulation of the glpFK operon encoding glycerol difihsion facilitator and glycerol kinase of Escherichia coli K-12. The Journal of Biological Chemistry 267:6122-6131. Wiese, D. E., B. R. Ernsting, R. M. Blumenthal, and R. G. Matthews. 1997. A nucleoprotein activation complex between the leucine-responsive regulatory protein and DNA upstream of the gltBDF operon in Escherichia coli. Journal of Molecular Biology 270:152-168. Wong, K. K., and H. S. Kwan. 1992. Transcription of glpt of Escherichia-coli-K12 is regulated by anaerobiosis and fur. Ferns Microbiology Letters 94:15-18. Wu, L. F., and M. A. MandrandBerthelot. 1995. A family of homologous substrate-binding proteins with a broad range of substrate specificity and dissimilar biological fiinctions. Biochimie 77:744-750. Xi, H. L., B. L. Schneider, and L. Reitzer. 2000. Purine catabolism in Escherichia coli and function of xanthine dehydrogenase in purine salvage. Journal of Bacteriology 182:5332-5341. Yamada, M., and M. H. Saier. 1988. Positive and negative regulators for glucitol (Gut) operon expression in Escherichia-coli. Journal of Molecular Biology 203:569-583. Yamamoto, K., K. Hirao‘, T. Oshima, H. Aiba, R. lUtsumi, and A. Ishihama. 2005. Functional characterization in vitro of all two-component Signal transduction systems from Escherichia coli. Journal of Biological Chemistry 280:1448-1456. Zhang, X. S., R. Garcia-Contreras, and T. K. Wood. 2007. chR (BhsA) influences Escherichia coli biofilm formation through stress response and surface hydrophobicity. Journal of Bacteriology 189:3051-3062. 131 105. 106. 107. Zhang, Z. G., M. Aboulwafa, M. H. Smith, and M. H. Saier. 2003. The ascorbate transporter of Escherichia coli. Journal of Bacteriology 185:2243-2250. Zhao, G. H., P. Ceci, A. Ilari, L. Giangiacomo, T. M. Laue, E. Chiancone, and N. D. Chasteen. 2002. Iron and hydrogen peroxide detoxification properties of DNA-binding protein from starved cells - A ferritin-like DNA-binding protein of Escherichia coli. Journal of Biological Chemistry 277 :27689-27696. Zimmer, D. P., E. Soupene, H. L. Lee, V. F. Wendisch, A. B. Khodursky, B. J. Peter, R. A. Bender, and S. Kustu. 2000. Nitrogen regulatory protein ‘ C-controlled genes of Escherichia coli: Scavenging as a defense against nitrogen limitation. Proceedings of the National Academy of Sciences of the United States of America 97: 14674-14679. 132 CHAPTER 5 COMPARATIVE ANALYSIS OF DIFFERENTIALLY EXPRESSED GENES IN EXIGUOBA C TERI UM SIBIRIC UM FOLLOWING EXPOSURE TO OXIDE NPS 133 Abstract NPs are being developed for diverse applications, such as water purification, drug delivery and electronic manufacturing. However, their environmental impact and toxicological implications are still unknown. Although numerous studies are now available about their antibacterial activity, information related to their impact at the level of gene expression are only beginning to emerge. In this study, microarrays were used to study the influence of Si02, Ti02, and ZnO NP on gene expression of E. sibiricum. The toxicity of NPS was consistent with their impact of gene expression. Both the number of repressed and induced genes generally increased in the order: Si02 to TiOz to ZnO. For all the three treatments, more genes were down-regulated genes compared to the number of up-regulated genes. The results demonstrated that they influenced expression of genes in different functional categories, including transportation, carbohydrate metabolism, lipid metabolism and DNA repair. Microbial respiration was suppressed during exposure of NPS, especially for aerobic respiration, as well as TCA cycle, oxidation of pyruvate and oxidative phosphorylation. Up-regulation of genes related to DNA damage repair and lipid or fatty acid biosynthesis, indicated that ROS induction in E. sibiricum was a response to NP exposure. 134 Introduction Nanotechnology is defined as “the understanding and control of matter at dimensions of roughly 1 to 100 nanometers, where unique phenomena enable novel applications” by National Nanotechnology Initiative. It is an emerging science that is revolutionizing the advancements taking place in many industries e.g., electronics, food, cosmetiCs, and medicines (http://www.nano.gov/html/facts/whatlsNano.htm1). It is estimated that by 2015 the market about nanotechnology will be around $1 trillion (44). Due to the increasing production and widespread application of diverse NMs, the concerns about the exposure to human and environmental release are also widespread (19, 59). However, studies related to the impacts of NMs on environment and human health are limited. Their small size, large surface area, shape, chemical composition, charge, solubility and aggregation, increase the difficulty to study and characterize their toxicological properties. NPS, the building blocks of NMs, have been used in diverse fields, such as environmental remediation, paint manufacturing and drug delivery. It has been reported that NPS had adverse impacts on bacteria and animals (16, 21, 31, 32, 58). However, there is little explanation about the mechanism of NP toxicity. It is hypothesized that the toxicity of NPS might be due to ROS induction (8, 44, 46, 49). DNA microarrays consist of short oligonucleotides or PCR products attached to a solid substrate. Oligonucleotide microchips containing multiple oligonucleotides are spotted on the chip surface. Probes targeting an entire microbial genome can be easily 135 represented in a Single array, making it feasible to perform genome-wide analysis. Since its inception more than a decade ago, DNA microchips have been used widely for gene expression study (6, 11, 51). DNA samples for analysis are labeled with fluorescent dyes and hybridized with the Oligonucleotide spots on the chip. The fluorescence pattern is then recorded by a laser scanner, and then Signals are quantified and analyzed using dedicated software to handle the amount of data. This technique can be extremely helpful in increasing our understanding about the effects of NP exposure to microorganisms and aiding in explaining the possible mechanisms of NP toxicity. Exiguobact‘erium sibiricum Strain 255-15, one member of the Bacillaceae, is a low G+C, Gram-positive, rod-shaped and non-spore-forrning psychrotrophic bacterium (53, i 62). It is wide-spread capable of living over a broad temperature range (-5 °C-40 °C) (54). The goal of this study was to explore the microbial response to different oxide NPS and possible mechanisms for their toxicity. More specifically, we wanted to evaluate the gene expression profiles of E. sibiricum when exposed to Si02, Ti02 and ZnO NPS. Moreover, we wanted to evaluate the influences of NPS on selected pathways focusing on the ROS damage response mechanisms in E. sibiricum. Materials and Methods Reagents and preparation of NP suspensions. Random hexamer primers, 100 mM nucleotide set, RNaseOUT, salmon sperm DNA, Superscriptase II RT, formamide, 10% sodium dodecyl sulfate, 20X Saline-Sodium Citrate (SSC), and 0.5 M 136 ethylenediaminetetraacetic acid (EDTA) were obtained from Invitrogen Co. (Carlsbad, CA). Bovine serum albumin (BSA), dimethyl sulfoxide, sodium hydroxide, 1 M hydrochloric acid and isopropanol were purchased from Sigma-Aldrich Chemical Co. (St. Louis, MO). RNAlater, SlideHyb buffer and aminoallyl-modified dUTP were obtained from Ambion Inc. (Austin, TX). Lysozyme, DNase I, Tris-Cl EDTA (TE) buffer, RNeasy mini kit, Quiaquick PCR purification kit and RNA protect bacteria reagents were obtained from Qiagen Co (Valencia, CA). Si02 (10 nm), Ti02 (5 nm) and ZnO (60 nm) were purchased from Sigma-Aldrich (St. Louis, MO). The 100 g/L NPs stock solutions were prepared in deionized and distilled water. Phosphate wash buffer (100 ml) for cDNA purification was made of 0.5 ml KzHPO4 (1 M, pH 8.5), 84.25 ml ethanol and 15.25 ml H20 and phosphate elution buffer (100 ml) contains 0.4 ml KZHPO4 (l M, pH 8.5) and 99.6 ml H20. All the solutions were autoclaved and stored at room temperature Cell growth, harvest and RNA extraction. From a freezing stock of E. sibiricum, streak plates were prepared with TSA medium and grown at 30 °C overnight. One loop of the plate was transferred to the tubes with 5 ml TSB. The NP stock solution was added to each tube with 100 ml TSB to result in final NP concentrations of 5000 ppm for Si02 NPs, 1000 ppm for Ti02 NPs and 500 ppm for ZnO NPS. The samples were inoculated and incubated until mid-log growth phase of E. sibicirum (0.1< g at 4 °C, and the supernatant was decanted. The cells were resuspended 137 in 1 ml RNAlater, transferred to a microcentrifuge tube and pelleted at SOOOX g at 4 °C for 10 min. The 100 pl lysozyme in TE buffer (3 mg/ml) was added to the cells, mixed by vortexing for 10 s, and incubated at room temperature for 30 min on a shaker. Buffer RLT (350 pl) in RNeasy mini kit was applied and vortexed vigorously for 30 5. After 2 min centrifiagation at maximum speed and removal of the sediment of NPS, the supernatant was mixed with 250 p1 ethanol by shaking vigorously (but not centrifuging after shaking). As described in RNeasy mini kit, the lysate was transferred to RNeasy Mini Spin column, and centrifuged for 14 s at >8000x g. DNase digestion was applied to the column by buffer RWl washing, DNase I solution washing (10 pl) incubating, and then buffer RWl re-washing. The RNA was purified as the protocol of RNeasy mini kit and the final concentration of RNA was measured by Nanodrop from Thermo Scientific Co. (Wilmington, DE). RNA from control samples was processed in a Similar way as NP samples. All experimental and control samples were prepared in triplicate. cDNA synthesis and dye-labeling. For each sample, 5 pg of total RNA was used to synthesize cDNA and two RNA populations of each sample were prepared for dye-swapping. RNA mixture contained 5 pg RNA, 5 pl random hexamer primers (3 mM), Rnase-free water in the final volume of 16.8 pl. The RNA mixture was incubated at 70 °C for 10 min and then cooled on ice for another 10 min. The RT—PCR reaction had a 30 pl volume and contained 6 pl 5X first strand synthesis buffer, 3 pl DTT (0.1 M), 1 pl RNaseOUT, 1.2 pl 25x dNTPs (7.5 mM aa-dUTP, 5 mM dTTP, 12.5 mM dCTP, 12.5 mM dGTP and 12.5 mM dATP), 2 pl Superscript 11 RT and 16.8 pl RNA mixture. The 138 RT-PCR reaction was incubated at 42 °C overnight, and then 10 pl NaOH (1 M) and 10 pl EDTA (0.5 M) were added to the reaction. After incubating at 65 °C for 15 min, 10 pl HCl was added to neutralize the pH. The cDNA was purified according to the protocol of Qiagen purification kit with some modification. The cDNA mixture was mixed with 300 pl PB buffer, and transferred to a Quiaquick column. After 1 min centrifuging at 17,900X g, the collection tube was emptied and the column was washed by 750 pl phosphate wash buffer twice and centrifuged for 1 min as above. After 1 min additional centrifuging, the column was transferred to a new tube, and then eluted twice by incubating the column with 30 pl phosphate elution buffer for 1 min and centrifuging for 1 min at maximum Speed. The cDNA sample Was dried using Speedvac from Global Medical Instrumentation Inc. (Ramsey, MN). The aa-dUTP cDNAwaS resuspended in 4.5 pl sodium carbonate (0.1 M, pH 9.0) and mixed with 4.5 pl Cy5 or Cy3. After incubating 2 h at room temperature in the dark, thirty-five pl NaOAc (100 mM, pH 5.2) was added and the dye-coupled cDNA was purified as described in the protocol of Qiaquick purification kit. The concentrations of cDNA and dye were measured by Nanodrop from Thermo Scientific Co. (Wilmington, DE). Hybridization. The glass-slide arrays were cross-linked at 6000 p] UV and waterbath was prewarmed to 46 °C. The slides were placed in prehybridization buffer containing 50% formamide, 5>< SSC, 0.1% SDS and 0.1 mg/ml BSA at 46 °C for 60 min. After washing three times with deionized and distilled water and once with isopropanol, the slides were dried by centrifuging at 1600 rpm for 3 min and transferred to dry and 139 clean tubes covered with foil. The coverslips were washed with RNase-free water and ethanol, dried by centrifuging 600 rpm and then placed in clean tubes covered with foil. The heat block was preheated to 95 °C and the hybridization oven to 60 °C. The labeled DNA of the control sample was resuspended in 35 pl hybridization buffer containing 50% formamide, 5>< SSC, 0.1% SDS, 0.1 pg/pl salmon sperm DNA. The other differently labeled DNA from NP-treated sample was suspended in the same hybridization solution, which was heated at 95 °C for 5 min. The dry slide-array was placed in a clean and dry hybridization chamber and coverSlip was placed on the slide. The hybridization solution was loaded on the slide, and 20 pl 5>< SSC was added in each well in the hybridization chamber. The chamber was placed in waterbath at 46 °C for 16 h. For each NP treatment, Six hybridizations from three biological replicates and two technical replicates (dye-swapping) were performed. Washing, scanning and data analysis. After 16 h, the chamber was removed from waterbath and dried with paper towel before opening. Wash Buffer l and slide staining jars were prewarmed at 46 °C. The Slide was taken out with forceps, and the coverslip was removed by immersing the array in a tube containing Wash Buffer 1 (1>< SSC and 0.1% SDS). The Slide was washed twice with Wash Buffer 1 (l X SSC and 0.1% SDS) for 5 min at 50 °C, then twice with Wash Buffer 2 (O.1>< SSC and 0.1% SDS) for 10 min at room temperature, and four times with Wash Buffer 3 (0.1>< SSC) for 1 min. The slide was dried by centrifuging at 1600 rpm for 3 min and transferred to dry tubes. Axon Genepix 4000B laser scanner was used for Slide scanning and Genepix 3.0 140 software was used to produce a raw data set. The scanner was turned on and warmed up for 15 min while the computer was off. The slide was placed face-down, and PMT and power was adjusted to yield optimal signals (as verified by histogram). A 5 pm pixels resolution setting was chosen for scanning. The area of interest was scanned and the image was saved. GENESPRING 6.0 software (Silicon Genetics, Redwood City, CA) was used for data analysis. The data was normalized both per chip and per gene (Lowess method), and dye-swapping. Only the spots with more than 55% of pixel greater than background plus 2X standard deviation in either Cy5 or Cy3 channel were used for analysis. Genes that showed a statistically significant change in gene expression (p>0.05) and a greater than or equal to 2-fold change in magnitude were regarded as significant. Results Global gene expression profiles under NP exposures. For all three NP exposures, the number of expressed genes exceeded the number of induced genes. The expression response to 500 ppm ZnO NPS was broader with more than 27% of the genome expressing differentially. For 1000 ppm Ti02 NPS, 18% of genome showed differential expression response while for 5000 ppm Si02 NPs, it was 14%. According to our previous viability data, it could be concluded that the exposure to more toxic NPS caused more differential expression in E. sibiricum. Among the up-regulated genes, almost 40% of Ti02-induced genes were also up-regulated in ZnO treatment, while only 15% of Ti02-induced genes were up-regulated in Si02 treatment (Figure 5.1(a)). Meanwhile, 141 48% of Ti02-repressed genes were down-regulated in ZnO treatment, whereas 27% of Ti02-repressed genes were down-regulated in Si02 treatment (Figure 5.1(b)). Thus the response of E. sibiricum to ZnO NP and Ti02 NP exposures was closer, compared to SiOz NP exposure. A total of 13 genes were induced in all the three treatment, and 38 genes were repressed by all three different NP exposures in Figure 5.1. Gene expression profiles following Si02 treatment. For Si02 treatment, fold-expression for 2539 genes ranged between 0.01 and 11.1. Based on the available information in the databases of Kyoto Encyclopedia of Genes and Genomes, UniProt and InterPro, the set of 234 down-regulated genes in the response to Si02 NPS were grouped into 17 functional categories. Categories “transportation” (24.8%) and “hypothetical” (20.9%) were the two dominant categories. Others included “biosynthesis” (8.5%), “carbohydrate metabolism” (7.3%), “amino acid metabolism” (6.4%), “transcription” (5.6%), “nucleotide metabolism” (3.8%), and “energy metabolism” (3.8%). Similarly, 158 up-regulated genes were grouped into the same 17 functional categories. Besides “hypothetical” (20.9%) and “transportation” (12.7%), most of induced genes belonged to the categories of “unknown” (9.5%), “carbohydrate metabolism” (9.5%), “transcription” (7.0%), “replication and repair” (7.0%), “amino acid metabolism” (6.3%), “translation” (5.1%), “lipid metabolism” (4.4%), “biosynthesis” (4.4%) and “cellular process” (3.8%). There were more down-regulated genes than up-regulated ones in categories of “biosynthesis” and “xenobiotics degradation and metabolism”, which might mean that 142 the bacteria closed some unnecessary pathways to save energy for survival. Meanwhile, more genes were induced compared to repressed ones in “lipid metabolism”, “replication and repair” and “signal transduction”, which could be related to the damage to membrane and DNA by ROS induction. Response of “transportation” related to genes in Si02 NP exposure. Transportation and internalization of NPs is an important factor influencing the fate of NPS. Because NPs are generally much larger than all the transportation channels, the mechanisms about how they enter the bacterial cells require additional exploration. In all the expression profiles, “transportation” was always one of the largest groups for both of induced and repressed genes. During the exposure to SiOz NPS, there were 58 down-regulated genes and 20 up-regulated genes in the “tranSportation” category. Among the repressed genes, 41.4% belonged to ABC transporters, 6.9% related to Na+ symporters, 5.2% functioned as PTS components, and 13.8% involved in sugar or amino acid transportation. For the induced genes in “transportation”, 50.0% were ABC-transporters, 10.0% K+ transporters and 5.0% were involved in PTS system. ABC transporters are a family of membrane proteins that function in the ATP-dependent transportation of many compounds across cellular membranes (1, 23). ABC transporters were found in both induced and repressed genes. The down-regulated genes mainly involve in transporting sugar, amino acid and peptides, such as ABC-type sugar transport system permease components and olio-peptide ABC transporter permease. 143 The up-regulated ones are related to ferric and other metal ions, such as ABC-type Fe3+-hydroxamate transport system components and ferric anguibactin transport system permease. Zhang et al. also reported over-expression of ABC transporters in mouse cells by poly(ethylene glycol)-block-polylactide NPs (66). ROS is known to oxidize fatty acids of phospholipids on membrane and cause lipid peroxidation (2). This process can change the membrane permeability and fluidity, and interrupt nutrient uptake, which may explain the response of amino acid and carbohydrate transport genes. Iron ion is another damage target for ROS, which may change the expression of metal ion ABC-transporter. Sodium/solute symporters and Potassium (K+) transporters were the other two groups of transporters influenced by Si02 NPS. Sodium/solute symporters catalyze the uptake various solutes into the microbial cells (50). Under extreme conditions, bacterial cells can use Na+ as coupling ions to generate Na+-electrochemical gradients by Na+ pumps and NaIL/HJr antiporters, and this is important for accumulation of nutrients (25, 26). K+ ion, as a major intracellular cation, plays a fundamental role for many physiological activities, such as protein synthesis and enzyme activation (14, 29). Under some conditions, Na+ substitutes functionally for K+ (55). In the Si02 treatment, Na+ symporters were repressed while K+ transporters were induced. The rapid leakage of K ions in bacterial cells during NP exposure was reported by Saito et al., and this may be related to the up-regulation of these K+ transporters influenced by Si02 NPS (56). 144 Category of “carbohydrate metabolism” in Si02 treatment. “Carbohydrate metabolism” is another important category in the gene expression profile. It involves various biochemical processes and is helpful to understand how bacteria survive through the exposure to NPS. There were 17 repressed genes and 14 induced genes in this category. Among the down-regulated genes, 23.5% of the repressed genes were related to CoA, 23.5% to pyruvate, and 23.5% to glycosidases or glucosidases. CoA functions as an acyl group carrier and plays a fundamental role in TCA cycle as well as synthesis and oxidation of fatty acids. It was reported that NPS could cause oxidation of CoA, and decrease of CoA may influence the activity of COA-related enzymes (37, 43). CoA is also involved in the oxidation of pyruvate, which may explain why the genes for pyruvate dehydrogenases were also down-regulated. The pyruvate dehydrogenase complex transforms pyruvate into acetyl-CoA, controls pyruvate entry into TCA cycle, and also participates in glycolysis and glucose oxidation (38). It may be concluded that Si02 NPS can repress the TCA cycle by oxidizing CoA and down-regulating the pyruvate dehydrogenases. The Si02 NPs also down-regulated the genes encoding glucosidases, glucosidase-related glycosidases and l-phosphofructoknase. These enzymes are involved in glycolysis, generating ATP and producing pyruvate as a part of aerobic respiration. For up-regulated genes in “carbohydrate metabolism”, there was no clear pattern among them. However, more than half of them were dehydrogenases, including alcohol dehydrogenases and aldehyde dehydrogenases, which transfer protons or electrons to 145 acceptors in electron transport chains, usually some coenzymes such as nicotinamide adenine dinucleotide/ nicotinamide adenine dinucleotide phosphate (NAD/NADP) (60). It is well known that NAD/NADP participates in generation of ATP, which might explain how E. sibiricum survives the exposure to Si02 NPs. Most of dehydrogenases contain sulfur-iron cluster, which is another attacking target for ROS. There were some coenzyme-dependent enzymes that were up-regulated in this category, such as flavin-dependent oxidoreductases, which may also adjust the cellular redox conditions since NPS can induce intracellular ROS. Category of “energy metabolism” in Si02 treatment. “Energy metabolism” is important for understanding how E. sibiricum produces energy and survives in the exposure to Si02 NPs. Almost 80% of repressed genes in this category were the subunits of cytochrome c oxidase or cytochrome caa3 oxidase. They function in one of electron transport chains in oxidative phosphorylation, and this process is important for cellular metabolism and energy synthesis as a part of aerobic respiration. Many enzymes involved in oxidative phosphorylation are located on membrane, which is one of the main attacking targets for ROS. ROS can also be produced during oxidative phosphorylation, such as free radicals and hydrogen peroxide, which cause some cellular damage. Among the induced genes in this category, it seems most could serve as antioxidant preventing the damages from ROS. For example, one gene encoding a catalase which can catalyze and converse hydrogen peroxide into water and oxygen, was up-regulated (24). 146 Two genes related to cysteine synthesis, which is beneficial for glutathione system in detoxification metabolism, were also up-regulated (40, 41). This could be another evidence for ROS damage, since cysteinyl residues in proteins are oxidized at the presence of ROS (22). Category of “metabolism of cofactors and vitamins” in $102 treatment. It is well known that cofactors and vitamins involve many biological processes and own diverse biochemical functions, such as activation of apoenzymes and regulation of microbial growth. Some of them function as antioxidant preventing microbes from ROS damages. Factors related to ROS damage are important in determining the response to NPS. For SiOz NP treatment, there were 6 repressed genes and 2 induced genes. Among the down-regulated genes, half were coenzyme F390 synthetafses. Coenzyme F390 system was first found in Methanobacterium spp., and this system can sense the intracellular reduction and oxidation potential (61). Coenzyme F390 usually is present when the cellular redox potential increases. 80 it may function as a signal compound, transferring information about intracellular redox potential. Two genes encoding ketopantoate reductase and thiamine monophosphate synthase respectively were repressed. The ketopantoate reductase is involved in biosynthesis of pantothenate, which is needed to form CoA and important for carbohydrate metabolism. The thiamine monophosphate synthase participates in thiamine metabolism, which relates to modification of CoA. It 147 seems that oxidation of CoA by NPs influences the relative pathways to other cofactors and vitamins. In the Si02 treatment, a gene encoding isochorismate synthase was up-regulated. Isochorismate synthase participates in ubiquinone biosynthesis and ubiquinone, also known as coenzyme Q (CoQ), usually functions as an antioxidant. So the induction of isochorismate synthase probably benefits the detoxification and protection from oxidative damages. Another induced gene was related to nitroreductase, which is involved in the reduction of nitrogen-containing compounds, such as NADH or FMN. Category of “replication and repair” in Si02 treatment. Since the induction of ROS by NPs and their genotoxicity have been reported before, “replication and repair” category may be important in understanding how E. sibiricum protects itself from DNA damage by NPS. For Si02 NP treatment, there were 5 repressed genes and 11 induced genes in this category. Thus more DNA repair systems were activated to alleviate DNA damages. Two bacteriophage integrases and one UV-damage repair protein were down-regulated by Si02 NPS. It is well known that Short-length UV can induce prophage-related genes. Therefore it appears that UV-related DNA repair system was repressed and not applicable for the DNA damage by ROS and NPs. The UV-damage repair protein was repressed in all the three treatments. It was reported that UV-irradiated TiOz had been applied as a photocatalyst for bacterial killing, and the suppression of UV-damage repair protein by NPS could worsen the bacterial survival during the 148 photo-killing (34, 52, 64). The induced genes included transposases, recombinases, ribonucleases, excisionases and oxidative damage protectants. Almost half of these genes were related to recombination, which may be important for DNA repair process. Gene expression profiles following Ti02 treatment. A total of 2681 genes showed a fold-change in the expression ranging from 0.01 to 17.31. As illustrated in Figure 5.2, the 289 down-regulated genes in the Ti02 treatment were grouped into 17 functional categories, including “hypothetical” (25.3%), “transportation” (15.2%), “carbohydrate metabolism” (8.7%), “amino acid metabolism” (8.7%), “transcription” (7.3%), “unknown” (5.2%) and “repair and replication” (4.2%). Similarly, 197 up-regulated genes were grouped into the same 17 functional categories. Besides “transportation” (26.4%) and “hypothetical” (23.4%), most of induced genes belonged to the categories of “unknown” (5.6%), “replication and repair” (5.6%), “transcription” (5.1%), “lipid metabolism” (4.6%), “carbohydrate metabolism” (4.6%), and “nucleotide metabolism” (4.1%). Usually there were both induced and repressed genes in the same functional categories, whereas more genes were repressed in the categories of “amino acid metabolism , carbohydrate metabolism , transcription” and “translation”, which may mean that E. sibiricum survived exposure to Ti02 NPS by reducing cellular activity. In “carbohydrate metabolism”, there were 25 down-regulated genes and 9 up-regulated genes. Among the repressed genes, 5 genes were related to CoA and another 5 genes were alcohol or pyruvate dehydrogenases. In the category of “transportation”, a 149 total of 44 genes were down-regulated and 52 genes were up-regulated. The repressed genes included ABC-transporters (43.2%), PTS components (9.1%), and Na+ or H+ symporters (4.5%), while the up-regulated genes covered ABC-transporters (42.3%), ferrous related transporters (11.5%), PST components (5.8%), antiporters or sympoters (5.8%), and K+ or Na+ transporters (3.8%). Similarly to Si02 treatment, Ti02 NPs repressed a few cytochrome c oxidases and induced some genes encoding thioredoxin-like proteins or NADPquuinone reductases in the “energy metabolism” category. It is well known that thioredoxin is another enzyme system functioning as an antioxidant and it can prevent E. sibiricum from oxidative damage by NP exposure. The NADPquuinone reductases participates in oxidative stress resistance by acting NADPH with a quinine as an acceptor (63). In the “replication and repair” category of Ti02 treatment, some UV-related repair genes were down-regulated, such as UV-damage repair protein and DNA photolyase. Recombinases and recombinational DNA repair proteins (23.1%) were also repressed by Ti02 NPs. Helicases (18.2%), transposases (18.2%) and DNA-directed RNA polymerases (18.2%) was among the induced genes by Ti02 NPS. Gene expression profiles following ZnO treatment. In the ZnO treatment, globally the fold-change in expression of 2616 genes ranged from 0.01 to 19.9. There were 402 down-regulated genes and 365 up-regulate genes. ZnO NPs had the largest number of up- or down-regulated genes (compared to in Si02 and Ti02 treatments). The down-regulated 150 genes can be grouped into 16 functional categories, including “hypothetical” (22.4%), “transportation” (13.4%), “carbohydrate metabolism” (12.2%), “amino acid metabolism” (9.5%), “transcription” (8.5%), “unknown” (6.7%) and “repair and replication” (4.0%) (Figure 6.2). Similarly, the up-regulated genes were grouped into the 17 functional categories, including “hypothetical” (18.4%), “transportation” (14.8%), “translation” (11.0%), “replication and repair” (9.3%), “unknown” (5.8%), “nucleotide metabolism” (5.8%), “amino acid metabolism” (5.8%), “carbohydrate metabolism” (5.2%) and “lipid metabolism” (4.4%). In ZnO treatment, more genes were repressed in the categories of “amino acid metabolism”, “carbohydrate metabolism”, and “transcription”, while more genes were induced in the categories of “replication and repair”, “lipid metabolism” and “nucleotide metabolism”. It can be concluded that E. sibiricum reduced cellular activity to save energy and survive the exposure to ZnO NPs, and activated relative pathways to repair DNA damages by ROS induction. There were 49 repressed genes and 19 induced genes in “carbohydrate metabolism”. Similar to the treatment of Si02 and Ti02 NPS, it was found that 5 repressed genes were related to CoA and another 10 repressed genes were alcohol or pyruvate dehydrogenases. In the category of “transportation”, there were 54 repressed genes and 54 induced genes. ABC-transporters (38.9%), PTS components (9.3%), Na+ or K+ transporters (9.3%), and symporters (5.6%) were found in the repressed transportation genes while there were ABC-transporters (38.9%), ferrous and ferric transporters (9.3%), antiporters (5 .6%), K+ 151 or NaJr transporters (5.6%), and PST components (1.9%) among the induced genes. ZnO NPS repressed a few cytochrome c oxidases, while they also induced some coenzyme related oxidoreductases. In ZnO treatment, there were 16 repressed genes and 34 induced genes in “replication and repair” category. Thus more DNA repair mechanisms were activated to alleviate the DNA damages from ZnO NPS. Among the down-regulated genes, helicases (12.5%), DNA-directed RNA polymerases (12.5%), transposases (25%) and recombinases or recombination proteins (12.5%) contributed the majority. The up-regulated genes included helicases (8.8%), DNA-directed RNA polymerases (11.8%), transposases (5.9%) and recombinases or recombination proteins (8.8%), DNA polymerases (8.8%), nucleases (11.8%), DNA repair proteins (8.8%) and topoisomerases (5.9%). Category of “metabolism of cofactors and vitamins” in Ti02 and ZnO treatments. There were 9 repressed genes and 6 induced genes by Ti02 NPs, while there were 12 repressed genes and 11 induced genes in the ZnO treatment. Similar to Si02 treatment, coenzyme F390 synthetase was down-regulated, which demonstrated the changes of intracellular redox potential. Dihydroneopterin aldolase, which participates in folate biosynthesis, was down-regulated in both Ti02 and ZnO NP treatments. It is well known that folate is important and necessary for DNA replication and cell division (27). Glutamyl-tRNA reductase was another repressed gene in both of treatments, which participates in porphyrin metabolism. Porphyrin is synthesized by reaction of glycine and 152 succinyl-CoA from TCA cycle. It is likely that oxidation of CoA reduced the synthesis of porphyrin and repressed the activity of glutamyl-tRNA reductase. Porphyrin is contained in heme, which is one subunit of cytochrome c oxides. In Ti02 treatment, uroporphyrinogen decarboxylase, also related to heme biosynthesis, was down-regulated (57). In ZnO treatment, the genes related to metabolism of CoQ and vitamin B6 were repressed. Vitamin B6 assists in the balancing between Na+ and K+. The repression of vitamin B6 biosynthesis protein may be related to the rapid leakage of KL In both TiOz and ZnO NP treatments, thiamine pyrophosphokinase, which catalyzes the reaction to produce thiamin pyrophosphate, was induced. Glycine oxidase, which participates in thiamine metabolism, was induced too. Meanwhile, two genes related to molybdopterin biosynthesis were up-regulated. Molybdopterin usually functions as a cofactor necessary for sulfite oxidases and nitrate reductases, which are very important in bacterial electron transfer chains and supply electron acceptors for anaerobic respiration (28). Category of “xenobiotics degradation and metabolism” in Si02, Ti02 and ZnO treatment. Although it is not a large category, it contained some interesting information about NP exposure, especially among the down-regulated genes. In SiOz treatment, one gene involved with ethyl tert-butyl ether (ETBE) degradation was down-regulated, which also happened in the treatments of Ti02 and ZnO NPS. ETBE has been used as a fuel oxygenate in gasoline and it is toxic and relatively non-degradable (39). Oxidation of 153 ETBE involves electron chain transfer by a cytochrome P450 monooxygenase (3, 13). The genes related to degradation of phenylacetate or phenylacetic acid were down-regulated, including some phenylacetate-COA oxygenase subunits that may produce some TCA cycle intermediates, and phenylacetyl-CoA ligase that activates phenylacetic acid to phenylacetyl-CoA (12, 45). In SiOz treatment 83.3% repressed genes were related to phenylacetate or phenylacetic acid, 50.0% in Ti02 treatment and 66.7% in ZnO treatment. Type I pullulanase and NADH-dependent nitro/flavin oxidoreductase were induced in both Ti02 and ZnO treatment. Type I pullulanase can specifically cleave the or-l, 6-glycosidic linkages in pullulan and branched or linear oligosaccharides (33). NADH-dependent nitro/flavin oxidoreductase is one type of nitroreductases, participating in reduction of nitrogen containing compounds (9). Shared down-regulated and up-regulated genes in Si02, Ti02 and ZnO treatments. As shown in Figure 5.3, there were 13 repressed genes and 38 induced genes that were common to all the three treatments. The categories “hypothetical” and “transportation” contained the largest number of common genes for both down-regulated and up-regulated genes. Among the down-regulated genes, the category “biosynthesis”, “carbohydrate metabolism” and “xenobiotics degradation and metabolism” were important. The induced genes in the “transportation” category were associated with ABC transporters for metal ions, such as ferric or zinc, while the most of repressed genes in the same category were sugar transporters, including ABC sugar transporters and PTS 154 components. In “biosynthesis” category, all the repressed genes were related to polysaccharide synthesis. Among the repressed genes of “carbohydrate metabolism”, nearly half were COA-related when the other half were pyruvate-related. In the “energy metabolism” category, the repressed genes mainly function in bacterial electron transfer chain. All the down-regulated genes in the “metabolism of cofactors and vitamins” were related to F390 coenzyme, which may be influenced by the changes of intracellular redox potential, and all the repressed genes in the “xenobiotics degradation and metabolism” involve in degradation of phenylacetic acid or ETBE. The patterns in these categories were useful for choosing molecular target candidates of NP toxicity. Discussion The microarray data was quite consistent with the studies reported in this dissertation focusing on viability and ROS. The order for the toxicity of NPs to E. sibiricum was ZnO > TiOz> Si02, same as ROS induction. The microarray results displayed that the E. sibiricum exposed to ZnO NPS possessed the most induced and repressed genes and the expression fold was the widest ranging from 0.01 to 19.9. Ti02 NPs possessed more influence on gene expression than Si02 NPs. According to 1-way ANOVA analysis, there were 1,677 genes expressed differently among these three treatments. The microarray data also indicated that NPS might suppress microbial respiration. A few genes coding for pyruvate dehydrogenase subunits were repressed by NPs. These 155 genes participate in oxidative decarboxylation of pyruvate, which links glycolysis and TCA cycle. The down-regulation of these genes probably reduced the synthesis of acetyl-CoA, which is very important for TCA cycle. Meanwhile, some genes encoding COA-related enzymes were down-regulated during the exposure to NPS. It has been reported that NPs could oxidize CoA (43). In this study, several cytochrome c oxidases were repressed by NPs. Cytochrome c oxidases are involved in oxidative phosphorylation, one of the key pathways for aerobic respiration. The essential enzymes in glycolysis were not influenced, although the exposure to NPs down-regulated a couple of genes in glycolysis, not requiring oxygen. It appeared that NPs influenced the aerobic respiration more than the anaerobic respiration. Similarly, fullerenes or its derivatives have also been found to inhibit E. coli respiration of glucose, and also hinder dioxygen consumption and respiratory chain (35-37). During the SiOz treatment, there were a few alcohol and aldehyde dehydrogenase genes that were up-regulated. These genes facilitate the conversion between alcohols and aldehydes and play an important role in fermentation. Quite a few dehydrogenases and Oxidoreductases were also induced by Ti02 NPS. These included NADPH: quinone reductase and ferredoxin oxidoreductase. For the ZnO NP exposure, however, some genes encoding important enzymes in the TCA cycle, including malate dehydrogenase, Succinyl-CoA synthetase beta subunit, isocitrate dehydrogenase and pyruvate dehydrogenase E2, were induced. Some alcohol and aldehyde dehydrogenases were repressed in ZnO NP treatment. 156 Previously we have demonstrated that ROS was induced in E. sibiricum by these NPS, and gene expression supported this observation. After exposure to NPS, most of the repressed ABC transporters were related to sugars, while induced ABC transporters were mostly related to metal ions, especially for iron ions. Iron is essential for most living bacteria functioning as a component for some important molecules, such as ribonucleotide reductase and cytochromes (30, 65). One possible explanation for the up-regulation of iron transporters is that ROS, induced by NPS, attacks sulfur-iron clusters and causes release of iron ions. It may also be due to the participation of iron in some electron transfer reactions since other cellular respiration pathways were suppressed. Additional evidence is the induction of enzyme in the categories of “lipid metabolism” and “replication and repair”, Since ROS can cause damage to membrane and DNA. Although a few enzymes in these two categories were down-regulated, the reason may be that they participated in the inactive pathway, and furthermore the number of induced genes was larger than that of the repressed, especially in the ZnO treatment. It was also observed that the expression of many oxidoreductases was influenced by the NPS, which might be related to the intracellular changes of redox potential caused by ROS. No induction of superoxide dismutases was found in E. sibiricum. The category “transportation” was always the most influenced category in all NP treatments. ABC transporters were the most influenced class of membrane transport proteins for various substances by using energy from ATP-binding and hydrolysis (4, 5, 157 18, 47). Some Kir transporters were up-regulated. Leakage of K+ occurs in bacteria in response to exposure to Ti02 NPS, which mayb be the reason for the induction of K+ transporters (56). K+ ion plays an important role in maintaining cellular viability, and the accumulation of K+ ions influences the rates of cellular growth, respiration and protein synthesis (10, 17, 20, 42). A few Na+/Solute symporters were found down-regulated while some Na+/H+ antiporters were up-regulated. Both Na+ ion and H+ ion are critical for cell bioenergetics and physiology (15, 48). The down-regulation of Na+/solute symporters and up-regulation of Na‘L/H+ antiporters should be related to H+ or Na+ electrochemical gradient, which is essential for solute uptake and energy storage. The proton gradient is important during NP exposure, perhaps due to the suppression of respiration by NPS, although Na+ ion is particularly used as coupling ions in extreme environments (7, 26). The categories “translation” and “signal transduction” were also important in all the three treatment. Most of the influenced genes in the category “translation” coded ribosomal proteins or enzymes related to tRNA. In the “signal transduction” category, the influenced genes coded histidine kinases. Not many genes were influenced in the category “biosynthesis”, and most repressed genes were related to polysaccharide biosynthesis. In other categories, such as “amino acid metabolism” and “transcription”, no specific pattern was observed. In conclusion, the microarray data demonstrated that the three types of NPS tested in this study influenced the gene expression of E. sibiricum. ZnO NPS were the most toxic 158 one. Exposure to ZnO NPS resulted in more up- and down-regulated genes than either Si02 or TiOz NPS. For all three treatments, the number of down-regulated genes was higher than the number of up-regulated genes. ROS induced by NPs caused damage to membrane, protein and DNA. The functional category “transportation” was the largest for both the induced and repressed genes in these three treatments. These included ABC transporters, PTS, exporters, and antiporters/symporters. More transporters for carbohydrate, amino acid and drugs were found in the down-regulated group than in the up-regulated group, while there were more metal ion transporters induced. For down-regulated genes, the categories “transcription”, “amino acid metabolism” and “carbohydrate metabolism” were other important ones, maybe due to the slowdown activity and growth caused by NP exposure. A number of COA-related genes, which function in TCA cycle, glycolysis, biosynthesis of fatty acids and amino acid metabolism, were also repressed. Genes related to DNA repair were induced, maybe due to genotoxicity caused by NPs. The microarray data demonstrated that NPs influenced the genes of E. sibiricum in a variety of functional categories. 159 Upregulated genes Ti02 8102 l 2110 Downregulated genes Ti 02 , b SiOz ZnO Figure 5.1 (a) The up-regulated gene profile in three different treatments. (b) The down-regulated gene profile in the three treatments. 160 400 - 300 1 200 - Gene numbers 100 - - Amino acid metabolism - Blosythesis Carbohydrate metabolism [:1 Cellular process - Energy metbolism - Hypothetical m Lipid metabolism Metabolism of cofactors and vitamins - Nucleotide metabolism - Protein fate - Replication and repair - Signal transduction - Transcription - Translation {.259 . '3‘“: ”‘51 : m J F" at“... E — Irma £2 .3" — “)1", jI-J Treatment Figure 5.2 The functional categories of the gene profiles in Si02 NP, Ti02 NP and ZnO NP exposures. 161 Shared genes among all the three treatments of Si02, Ti02 and ZnO NPs - shared downregulated genes 8 - l:l shared upregulated genes Gene number Functional categories Figure 5.3 The functional categories of the shared genes among Si02 NP, Ti02 NP and ZnO NP exposures. 162 References 10. Baldan, A., P. Tarr, R. Lee, and P. A. Edwards. 2006. ATP-binding cassette transporter G1 and lipid homeostasis. Current Opinion in Lipidology 17:227-232. Cabiscol, E., J. Tamarit, and J. Ros. 2000. Oxidative stress in bacteria and protein damage by reactive oxygen species. International Microbiology 3:3-8. Chauvaux, S., F. Chevalier, C. Le Dantec, F. Fayolle, I. Miras, F. Kunst, and P. Beguin. 2001. Cloning of a genetically unstable cytochrome P-450 gene cluster involved in degradation of the pollutant ethyl tert-butyl ether by Rhodococcus ruber. Journal of Bacteriology 183:6551-6557. Davidson, A. L., E. Dassa, C. Orelle, and J. Chen. 2008. Structure, function, and evolution of bacterial ATP-binding cassette systems. Microbiology and Molecular Biology Reviews 72:317-364. Davidson, A. L., and P. C. Maloney. 2007. ABC transporters: how small machines do a big job. Trends in Microbiology 15:448-455. DeRisi, J. L., V. R. Iyer, and P. 0. Brown. 1997. Exploring the metabolic and genetic control of gene expression on a genomic scale. Science 278:680-686. Dimroth, P. 1994. Bacterial sodium ion-coupled energetics. Antonie Van Leeuwenhoek International Journal of General and Molecular Microbiology 65:381-395. Donaldson, K., and C. L. Tran. 2002. Inflammation caused by particles and fibers. Inhalation Toxicology 14:5-27. Emptage, C. D., R. J. Knox, M. J. Danson, and D. W. Hough. 2009. Nitroreductase from Bacillus licheniforrnis: A stable enzyme for prodrug activation. Biochemical Pharmacology 77:21-29. Ennis, H. L. 1971. Role of potassium in regulation of polysome content and 163 11. 12. 13. 14. 15. 16. 17. 18. 19. protein synthesis in Escherichia-coli. Archives of Biochemistry and Biophysics 142:190-&. Fawcett, P., P. Eichenberger, R. Losick, and P. Youngman. 2000. The transcriptional profile of early to middle sporulation in Bacillus subtilis. Proceedings of the National Academy of Sciences of the United States of America 97 :8063-8068. Ferrandez, A., B. Minambres, B. Garcia, E. R. Olivera, J. M. Luengo, J. L. Garcia, and E. Diaz. 1998. Catabolism of phenylacetic acid in Escherichia coli - Characterization of a new aerobic hybrid pathway. Journal of Biological Chemistry 273:25974-25986. Ferreira, N. L., C. Malandain, and F. Fayolle-Guichard. 2006. Enzymes and genes involved in the aerobic biodegradation of methyl tert-butyl ether (MTBE). Applied Microbiology and Biotechnology 72:252-262. Gaber, R. F., C. A. Styles, and G. R. Fink. 1988. Trkl encodes a plasma-membrane protein required for high-affinity potassiurn-transport in Saccharomyces-cerevisiae. Molecular and Cellular Biology 8:2848-2859. Grinstein, S., and H. Wieczorek. 1994. Cation antiports of animal plasma-membranes. Journal of Experimental Biology 196:307-318. Gupta, A. K., and M. Gupta. 2005. Cytotoxicity suppression and cellular uptake enhancement of surface modified magnetic nanoparticles. Biomaterials 26:1565-1573. Harold, F. M., and Y. Kakinuma. 1985. Primary and secondary transport of _ cations in bacteria. Annals of the New York Academy of Sciences 456:375-383. HolIenstein, K., R. J. P. Dawson, and K. P. Locher. 2007. Structure and mechanism of ABC transporter proteins. Current Opinion in Structural Biology 17:412-418. Holsapple, M. P., W. H. Farland, T. D. Landry, N. A. Monteiro-Riviere, J. M. 164 20. 21. 22. 23. 24. 25. 26. 27. 28. Carter, N. J. Walker, and K. V. Thomas. 2005. Research strategies for safety evaluation of nanomaterials, part II: Toxicological and safety evaluation of nanomaterials, current challenges and data needs. Toxicological Sciences 88:12-17. Hueting, S., T. Delange, and D. W. Tempest. 1979. Energy requirement for maintenance of the transmembrane potassium gradient in Klebsiella-aerogenes thc-418 - continuous culture study. Archives of Microbiology 123:183-188. Hussain, S. M., K. L. Hess, J. M. Gearhart, K. T. Geiss, and J. J. Schlager. 2005. In vitro toxicity of nanoparticles in BRL 3A rat liver cells. Toxicology in Vitro 19:975-983. Imlay, J. A. 2003. Pathways of oxidative damage. Annual Review of Microbiology 57:395-418. Jessup, W., I. C. Gelissen, K. Gaus, and L. Kritharides. 2006. Roles of ATP binding cassette transporters Al and G1, scavenger receptor BI and membrane lipid domains in cholesterol export from macrophages. Current Opinion in Lipidology 17:247-257. 1 Johnson, F., and C. Giuiivi. 2005. Superoxide dismutases and their impact upon human health. Molecular aspects of medicine 16:340-352. Jung, H. 2002. The sodium/substrate symporter family: structural and functional features. F ebs Letters 529:73-77. Jung, H. 2001. Towards the molecular mechanism of Na+/solute symport in prokaryotes. Biochimica Et Biophysica Acta-Bioenergetics 1505: 131-143. Kamen, B. 1997. Folate and antifolate pharmacology. Seminars in Oncology 24:30-39. Kisker, C., H. Schindeiin, D. Baas, J. Retey, R. U. Meckenstock, and P. M. H. Kroneck. 1998. A structural comparison of molybdenum cofactor-containing enzymes. F ems Microbiology Reviews 22:503-521. 165 29. 30. 31. 32. 33. 34. 35. 36. 37. K0, C. H., and R. F. Gaber. 1991. Trkl and Trk2 encode structurally related K+ transporters in Saccharomyces-cerevisiae. Molecular and Cellular Biology 11:4266-4273. Koster, W. 2001. ABC transporter-mediated uptake of iron, siderophores, heme and vitamin B-12. Research in Microbiology 152:291-301. Lewinski, N., V. Colvin, and R. Drezek. 2008. Cytotoxicity of nanoparticles. Small 4:26-49. Lin, W. S., Y. W. Huang, X. D. Zhou, and Y. F. Ma. 2006. In vitro toxicity of silica nanoparticles in human lung cancer cells. Toxicology and Applied Pharmacology 217:252-259. Malle, D., T. Itoh, W. Hashimoto, K. Murata, S. Utsumi, and B. Mikami. 2006. Overexpression, purification and preliminary X-ray analysis of pullulanase from Bacillus subtilis strain 168. Acta Crystallographica Section F-Structural Biology and Crystallization Communications 62:381-384. Maness, P. C., S. Smolinski, D. M. Blake, Z. Huang, E. J. Wolfrum, and W. A. Jacoby. 1999. Bactericidal activity of photocatalytic Ti02 reaction: Toward an understanding of its killing mechanism. Applied and Environmental Microbiology 65:4094-4098. Mashino, T., D. Nishikawa, K. Takahashi, N. Usui, T. Yamori, M. Seki, T. Endo, and M. Mochizuki. 2003. Antibacterial and antiproliferative activity of cationic fullerene derivatives. Bioorganic & Medicinal Chemistry Letters 13:4395-4397. Mashino, T., K. Okuda, T. Hirota, M. Hirobe, T. Nagano, and M. Mochizuki. 1999. Inhibition of E-coli growth by fullerene derivatives and inhibition mechanism. Bioorganic & Medicinal Chemistry Letters 9:2959-2962. Matsunaga, T., R. Tomoda, T. Nakajima, and H. Wake. 1985. Photoelectrochemical sterilization of microbial-cells by semiconductor powders. Ferns Microbiology Letters 29:211-214. 166 38. 39. 40. 41. 42. 43. 44. 45. 46. 47. McFate, T., A. Mohyeidin, H. Lu, J. Thakar, J. Henriques, N. D. Haiim, H. Wu, M. J. Schell, T. M. Tsang, O. Teahan, S. Zhou, J. A. Califano, N. H. Jeoung, R. A. Harris, and A. Verma. 2008. Pyruvate dehydrogenase complex activity controls metabolic and malignant phenotype in cancer cells. Journal of Biological Chemistry 283:22700-22708. McGregor, D. 2007. Ethyl tertiary-butyl ether: A toxicological review. Critical Reviews in Toxicology 37:287-312. Meister, A. 1988. Glutathione metabolism and its selective modification. Journal of Biological Chemistry 263: 17205-17208. Meister, A., and M. E. Anderson. 1983. Glutathione. Annual Review of Biochemistry 52:711-760. Meury, J., and A. Robin. 1985. A high-affinity site for glutathione in the cytoplasm of Escherichia-coli and its possible role in potassium retention. European Journal of Biochemistry 148:113-118. Mitoraj, D., A. Janczyk, M. Strus, H. Kiscb, G. Stochel, P. B. Heczko, and W. Macyk. 2007. Visible light inactivation of bacteria and fungi by modified titanium dioxide. Photochemical & Photobiological Sciences 6:642-648. Nel, A., T. Xia, L. Madler, and N. Li. 2006. Toxic potential of materials at the nanolevel. Science 311:622-627. O'Leary, N. D., W. A. Duetz, A. D. W. Dobson, and K. E. O'Connor. 2002. Induction and repression of the sty operon in Pseudomonas putida CA-3 during growth on phenylacetic acid under organic and inorganic nutrient-limiting continuous culture conditions. F ems Microbiology Letters 208:263-268. Oberdorster, G., E. Oberdorster, and J. Oberdorster. 2005. Nanotoxicology: An emerging discipline evolving from studies of ultrafine particles. Environmental Health Perspectives 113:823-839. Oldham, M. L., A. L. Davidson, and J. Chen. 2008. Structural insights into 167 48. 49. 50. 51. 52. 53. 54. 55. 56. ABC transporter mechanism. Current Opinion in Structural Biology 18:726-733. .Padan, E., M. Venturi, Y. Gerchman, and N. Dover. 2001. Na+/I-I+ antiporters. Biochimica Et Biophysica’Acta-Bioenergetics 1505:144-157. Park, E. J., and K. Park. 2009. Oxidative stress and pro-inflammatory responses induced by silica nanoparticles in vivo and in vitro. Toxicology Letters 184:18-25. Reizer, J., A. Reizer, and M. H. Saier. 1994. A functional superfamily of sodium/solute symporters. Biochimica Et Biophysica Acta-Reviews on Biomembranes 1197: 133-166. Richmond, C. S., J. D. Glasner, R. Mau, H. F. Jin, and F. R. Blattner. 1999. Genome-wide expression profiling in Escherichia coli K-12. Nucleic Acids Research 27:3821-3835. ' Rincon, A. G., and C. Pulgarin. 2004. Bactericidal action of illuminated Ti02 on pure Escherichia coli and natural bacterial consortia: post-irradiation events in the dark and assessment of the effective disinfection time. Applied Catalysis B-Environmental 49:99-1 12. Rodrigues, D. F., N. Ivanova, Z. L. He, M. Huebner, J. Z. Zhou, and J. M. Tied je. 2008. Architecture of thermal adaptation in an Exiguobacterium sibiricum strain isolated from 3 million year old permafrost: A genome and transcriptome approach. Bmc Genomics 9:—. Rodrigues, D. F., and J. M. Tiedje. 2007. Multi-locus real-time PCR for quantitation of bacteria in the environment reveals Exiguobacterium to be prevalent in permafrost. Ferns Microbiology Ecology 59:489-499. Rodriguez-Navarro, A. 2000. Potassium transport in fungi and plants. Biochimica Et Biophysica Acta-Reviews on Biomembranes 1469:1-30. Saito, T., T. Iwase, J. Horie, and T. Morioka. 1992. Mode of photocatalytic bactericidal action of powdered semiconductor Ti02 on mutans Streptococci. Journal of Photochemistry and Photobiology B-Biology 14:369-379. 168 57. 58. 59. 60. 61. 62. 63. 64. 65. Silva, P. J., and M. J. Ramos. 2005. Density-functional study of mechanisms for the cofactor-free decarboxylation performed by uroporphyrinogen III decarboxylase. Journal of Physical Chemistry B 109:18195-18200. Thill, A., O. Zeyons, O..Spalia, F. Chauvat, J. Rose, M. Auffan, and A. M. Flank. 2006. Cytotoxicity of Ce02 nanoparticles for Escherichia coli. Physico-chemical insight of the cytotoxicity mechanism. Environmental Science & Technology 40:6151-6156. Tsuji, J. S., A. D. Maynard, P. C. Howard, J. T. James, C. W. Lam, D. B. Warheit, and A. B. Santamaria. 2006. Research strategies for safety evaluation of nanomaterials, part IV: Risk assessment of nanoparticles. Toxicological Sciences 89:42-50. Unden, G, and J. Bongaerts. 1997. Alternative respiratory pathways of Escherichia coli: Energetics and transcriptional regulation in response to electron acceptors. Biochimica Et Biophysica Acta-Bioenergetics 1320:217-234. Vermeij, P., R. J. van der Steen, J. T. Keltjens, G D. Vogels, and T. Leisinger. 1996. Coenzyme F390 synthetase from Methanobacterium thermoautotrophicum Marburg belongs to the superfamily of adenylate-forming enzymes. Journal of Bacteriology 178:505-510. Vishnivetskaya, T. A., R. Siletzky, N. Jefferies, J. M. Tiedje, and S. Kathariou. 2007. Effect of low temperature and culture media on the growth and freeze-thawing tolerance of Exiguobacterium strains. Cryobiology 54:234-240. Wang, G, and R. J. Maier. 2004. An NADPH quinone reductase of Helicobacter pylori plays an important role in oxidative stress resistance and host colonization. Infection and Immunity 72: 1391-1396. Wei, C., W. Y. Lin, Z. Zainal, N. E. Williams, K. Zhu, A. P. Kruzic, R. L. Smith, and K. Rajeshwar. 1994. Bactericidal activity of Ti02 photocatalyst in aqueous-media - toward a solar-assisted water disinfection system. Environmental Science & Technology 28:934-938. Weinberg, E. D. 1997. The Lactobacillus anomaly; Total iron abstinence. 169 66. Perspectives in Biology and Medicine 40:578-583. Zhang, Y. D., Z. Y. Hu, M. Y. Ye, Y. F. Pan, J. J. Chen, Y. L. Luo, Y. Q. Zhang, L. X. He, and J. W. Wang. 2007. Effect of poly(ethylene glycol)-block-polylactide nanoparticles on hepatic cells of mouse: Low cytotoxicity, but efilux of the nanoparticles by ATP-binding cassette transporters. European Journal of Pharmaceutics and Biopharmaceutics 66:268-280. 170 TY "'llllllllllllllllllflllllllllllllllllllllll“