Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 + 2061 word(s) 2061 2021-11-02 10:31:08 |
2 format corrected. + 261 word(s) 2322 2021-11-10 08:05:50 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Ouyang, S. Genotoxicity of Graphene Family Nanomaterials on DNA. Encyclopedia. Available online: https://encyclopedia.pub/entry/15850 (accessed on 27 July 2024).
Ouyang S. Genotoxicity of Graphene Family Nanomaterials on DNA. Encyclopedia. Available at: https://encyclopedia.pub/entry/15850. Accessed July 27, 2024.
Ouyang, Shaohu. "Genotoxicity of Graphene Family Nanomaterials on DNA" Encyclopedia, https://encyclopedia.pub/entry/15850 (accessed July 27, 2024).
Ouyang, S. (2021, November 10). Genotoxicity of Graphene Family Nanomaterials on DNA. In Encyclopedia. https://encyclopedia.pub/entry/15850
Ouyang, Shaohu. "Genotoxicity of Graphene Family Nanomaterials on DNA." Encyclopedia. Web. 10 November, 2021.
Genotoxicity of Graphene Family Nanomaterials on DNA
Edit

Graphene family nanomaterials (GFNs), including graphene, graphene oxide (GO), reduced graphene oxide (rGO), and graphene quantum dots (GQDs), have manifold potential applications, leading to the possibility of their release into environments and the exposure to humans and other organisms. However, the genotoxicity of GFNs on DNA remains largely unknown. In this review, we highlight the interactions between DNA and GFNs and summarize the mechanisms of genotoxicity induced by GFNs.

graphene family nanomaterials genotoxicity DNA damage safety toxicity

1. Introduction

Currently, GFNs, as promising nanomaterials, have attracted increasing attention in the scientific community and are in commercial production for many applications, such as energy storage [1][2][3][4][5][6][7][8], medicine [9][10][11][12][13][14][15][16], environmental protection [17][18][19][20][21][22], and industrial manufacturing [23][24][25]. For example, the market for graphene-based products is forecast to reach $675 million by 2020 [26]. With rapid developments in application and production of GFNs, their potential for release into the environment and the environmental risks of GFNs have become emerging issues [27][28][29]. Consequently, many studies have shown that adverse effects can be induced by GFNs in vivo and in vitro, such as organ (e.g., lung, liver, and spleen) toxicity, cytotoxicity, immunotoxicity, neurotoxicity, and reproductive and developmental toxicity [30][31]. Moreover, the toxicity mechanisms of GFNs to organisms, including physical destruction, oxidative stress, inflammatory response, apoptosis, autophagy, and necrosis, are summarized in Table 1 . However, the genotoxicity of GFNs on DNA (e.g., DNA damage) remains largely unknown.

Table 1. The toxicity of GFNs in vivo and in vitro.
Products Supplier or Synthesis Methods Dose Animal or Cell Models Toxicological Mechanisms Adverse Effects Ref.
graphene nanoplatelets cheaptubes.com (Brattleboro, VT, USA) 0.3, 1 mg/rat rat oxidative stress, inflammation lung inflammation [32]
commercial GO and rGO Nanjing XFNANO Materials Tech Co., Ltd., (China) 2.0 mg/kg body weight rat transcriptional and epigenetic liver zonated accumulation [33]
amination GQDs
carboxylated GQDs
hydroxylated GQDs
Nanjing XFNANO Materials Tech Co., Ltd., (China) 100, 200 μg/mL A549 cells autophagy cytotoxicity [34]
GO and rGO oxidated from carbon nanofibers Grupo Antolin
(Spain)
0.1, 1.0, 10, 50 mg/L erythrocyte cell oxidative stress genotoxicity [35]
GO nanosheets Sigma-Aldrich (St. Louis, MO, USA) 40, 60, 80 mg/L Human SH-SY5Y neuroblastoma cell oxidative stress, autophagy–lysosomal network dysfunction cytotoxicity [36]
pristine rGO Chengdu Organic Chemicals Co., Ltd., the Chinese Academy of Sciences 1–100 mg/L Earthworm coelomocytes oxidative stress immunotoxicity [37]
single layer GO
(product no. GNOP10A5)
ACS Materials LLC (Medford, MA, USA) 1, 10, 50, 150, 250, 500 mg/L Escherichia coli physical destruction toxicity against bacteria [38]
GO modified Hummers method 25 mg/L THP-1 and BEAS-2B cells lipid peroxidation, membrane adsorption, membrane damage cytotoxicity [39]
GO modified Hummers method 2 mg/kg rat lipid peroxidation, membrane adsorption, membrane damage acute lung inflammation [39]
GO Nanjing XFNANO Materials Tech Co., Ltd., (China) 0–100 mg/L zebrafish embryos oxidative stress developmental toxicity [40]
GO modified Hummers method 10 mg/L Caenorhabditis elegans oxidative stress toxicity [41]
graphene,
GO
modified Hummers method 3.125–200 mg/L human erythrocytes and skin fibroblasts oxidative stress cytotoxicity [42]
graphene exfoliated form graphite,
GO oxidated from carbon fibers
Grupo Antolin Ingeniería (Burgos, Spain) 1, 10 mg/L primary neurons inhibition of synaptic transmission, altered calcium homeostasis neurotoxicity [43]

Genotoxicity is broadly defined as ‘damage to the genome’ and also a distinct and important type of toxicity, as specific genotoxic events are considered hallmarks of cancer [44]. Generally, the genotoxicity can be sub-classified into direct genotoxicity and indirect genotoxicity in cells or the nucleus [45][46][47]. Nanoparticles (NPs) can be uptaken by the nucleus and induce DNA damage, leading to direct genotoxicity on organisms [46]. While many studies have shown that most NPs cannot enter the nucleus, they still indirectly affect genotoxicity by oxidative stress, epigenetic changes, inflammation, and autophagy [46]. Moreover, genotoxicity plays a key role in assessing the safety of NPs on human health and the environment [48][49][50][51]. Although there has been many researches about the genotoxicity of NPs in recent years, it is mainly focused on traditional artificial nanomaterials, such as TiO 2, carbon nanotubes, and silver and gold NPs [52][53][54]. However, the existing literature on genotoxicity of GFNs remains limited and conflicting. A few studies showed that GFNs had no adverse effects on genotoxicity [55]. In contrast, many researchers have reported that the small size and sharp edges of GFNs (e.g., GO and GQDs) can induce genotoxicity on aquatic organisms (e.g., fish and algae) [56][57][58]. However, the direct and indirect genotoxicity mechanisms of GFNs remain unclear, despite genotoxic phenomena being widely reported.

2. Factors Influencing Genotoxicity of GFNs

As is known to all, there is a strong correlation between cytotoxicity and the physicochemical properties of NPs, such as particle size and shape, surface characteristics, and surface functionalization. Similarly, the genotoxicity of GFNs can be affected by these factors [59]. The genotoxicity of GFNs is greatly varied in the literature, which can be attributed to numerous factors including physicochemical properties (morphology, surface chemistry, size, shape, and purity), dose, test species, exposure time, and exposure assay [60][61].

2.1. Surface Properties

The oxygen-containing functional groups play a key role in the genotoxicity of GFNs [35][62][63][64][65]. For example, the rGO with lower oxygen content can induce stronger genotoxicity on ARPE-19 cells than these GO with higher oxygen content, suggesting that GO has a better biocompatibility owing to more saturated C–O bonds [62]. The remove of epoxy groups from the GO surface mitigates GO in vivo genotoxicity toward Xenopus laevis tadpoles [35]. Compared with GO, graphene, rGO, and graphite all induce higher levels of genotoxicity in glioblastoma multiforme cells, and the difference was attributed to the hydrophilic and hydrophobic surface and edge structure of GFNs [65]. GO has hydrophilic properties and smooth and regular edges, while rGO and graphene have hydrophobic properties and sharp and irregular edges, which can damage the integrity of cell membranes greatly. The carboxyl groups in the surface of carboxyl-FLG may scavenge oxidative radical on bronchial epithelial cells to alleviate the genotoxicity of FLG [64]. Moreover, different immunological mechanisms triggered by GFNs can be attributed to the proportion of hydroxyl groups [63]. Cells produce a stronger inflammatory response after being exposed to GO than rGO by detecting transcriptomic changes, and the reason is attributed to the large number of hydroxyl groups on the surface of GO [63]. The surface functionalization also can significantly modulate the toxicity of GFNs [57][66][67][68]. For example, amino functionalized GQDs induced lower ferroptosis effects than nitrogen-doped GQDs [66]. Similarly, the DNA methylation of various tissues induced by GQDs was depend on their different surface chemical modifications [57]. Increased cytotoxicity and genotoxicity of the aminated GO were detected by following 24 h exposure on Colon 26 cells [68]. A study on the genotoxicity reduced by GO and rGO showed that the GTPs-rGO reduced by green tea polyphenols (GTPs) yielded more biocompatible and reduced sheets with lower genotoxic effects, as compared to the N2H4–rGO, which were reduced by hydrazine (N2H4) [69]. The acid-polyethylene glycol (LA-PEG) and PEG modified GO induced gentle DNA damage and decreased the genotoxicity of GO to HLF cells [67]. Surface charge also influences significantly the genotoxicity of GFNs [67][70]. The genotoxic effect of GO on cells is proportional to the amount of positive charge on the surface [67]. The surface charge density of graphene in aqueous solution can transform to chemically-converted graphene, leading to the capture of large amounts of DNA [70]. The different hydrophilic and hydrophobic properties of GO/rGO regulated by differential surface chemistry (especially the O/C ratio) determine the potential of graphene to interact with organisms [71][72][73]. Despite hydrophilic and hydrophobic rGO exhibiting similar toxic responses (e.g., cytotoxicity, DNA damage, and oxidative stress) to cells, their biological and molecular mechanisms are different [71]. The hydrophilic GO and hydrophobic rGO induce both kinds of DNA damage, namely single stranded and double stranded breaks, but the dose dependency was very significant and evident in GO exposure in DNA damage but not in rGO exposure [71]. Hydrophilicity, also an important factor in determining the biocompatibility and colloidal stability of GFNs, leads to different interactions with cells and bio-distribution of GFNs [72][73]. For example, simple accumulation of hydrophobic pristine graphene on the surface of monkey kidney cells without any cellular internalization led to severe metabolic toxicity, whereas hydrophilic GO was internalized by the cells and concentrated near the perinuclear region without causing any toxicity under lower concentrations [72]. Therefore, the surface properties play an important role in understanding the genotoxicity manifestations and biological and molecular mechanisms of GFNs.

2.2. Size and Structure

The genotoxicity of GFNs within organisms is size-dependent. Compared with large GFNs, small GFNs have bigger surface areas and provide more sites to interact with cells, leading to greater cellular uptake of GFNs [74]. The size effect plays a key role in the genotoxicity of GFNs. For example, small rGO (average lateral dimensions 114 nm) induce higher genotoxicity in the hMSCs than large rGO (3.8 ± 0.4 μm) at 0.1 and 1.0 μg/mL after 1 h exposure. The lateral size and extremely sharp edged structure of GFNs can result in higher permeability to the cell and nucleus, resulting in greater genotoxicity. Similarly, the size of GFNs is an important determinant of subcellular penetration [74]. Li et al. [75] suggested that the larger the lateral size of GO, the more severe is the pyroptosis induced by GO in Kupffer cells. Moreover, there is a strong correlation between the size of GO and the structural change in small-interfering RNAs [76]. The large GO merely reduces the A-helical pitch, while small GO inserted into the double strands can wreak havoc on the RNA conformation [77]. In addition, Kong et al. [78] proved that the DNA damage mechanism of GQDs was limited by the size of GQDs through molecular dynamics simulations. Briefly, the relatively large GQDs (61 benzene rings) tend to stick to the ends of the DNA molecule, causing the DNA to unfold, while the small GQDs (seven benzene rings) are easily embedded in DNA molecules, leading to DNA base mismatches. The planar structure of GFNs may also have an effect on DNA damage. The dsDNA bases have a stronger binding affinity with wrinkled GFNs and even cause more DNA damage than with planar GFNs [79]. Given these discordant results, it is necessary to clarify the size- and structure-related genotoxicity of GFNs.

2.3. Exposure Dose and Time

The dose–response relationship is an important principle in nanotoxicology [46]. The modified GQDs may induce DNA hypermethylation in a time and dose dependent manner [57]. The high-dose (50 mg/L) GO induces more serious DNA methylation (hypermethylation) than low-dose (10 mg/L) treatment [80]. The effective accumulation of GFNs in the nucleus is regulated by two nuclear pore complex genes (Kapβ2 and Nup98), and their cellular internalization and absorption are related to exposure time [81]. Notably, the rGO sheets with the same size or larger size, higher concentration (100 μg/mL), and longer exposure time (24 h) showed no obvious genotoxicity in the hMSCs [82]. Overall, there are few studies on the genotoxicity of GFNs doses, and especially the combination of GFNs type and dose exposure is rare.

2.4. The Resistance of Cell Structures and Biological Barriers

From an organism’s perspective, the responses of various types of cells, organs, and tissues with different structures and functions to GFNs exposure were highly diverse. Internalization and direct contact membrane stress with extremely sharp edges of GFNs are considered as important mechanisms of toxicity [83][84]. For different bacterial models to graphene toxicity, the outer membranes can better “protect” bacteria from graphene [85]. The biological barrier is crucial for mammals against the damage from GFNs [30][59]. Both GO and graphene were able to induce DNA breaks in an in vitro model simulating the human intestinal barrier [86]. Moreover, GO nanosheets could break through the first line of host defense by disrupting the ultrastructure and biophysical properties of lung surfactant membranes [87]. Combined with the routes and doses of human exposure, relevant biological barriers toxicity can be considered as an aspect of assessing GFNs genotoxicity.

3. Genotoxicity Testing of GFNs

3.1. Detection of GFNs in Cells and Organism Tissues

The detection of GFNs internalization (distribution and behavior) in model organisms and cells is a key step for a better understanding of their genotoxicity and underlying mechanisms. The most commonly used detection technique includes direct observation of localization of GFNs in organisms and cells by transmission electron microscopy (TEM) [88]. The hyperspectral imaging is also used to visualize cellular interactions with NPs [89], such as cellular uptake and binding of GFNs [90]. The label-based approaches to image GFNs exist in cells by confocal and fluorescence microscopy, reflection-based imaging, and flow cytometry. Additionally, scanning electron microscopy (SEM) can be used to detect the attachment of GFNs in the surface zone of cells [56][90]. Raman spectroscopy and atomic force microscopy (AFM) were used to evaluate nuclear area changes and the disruption of DNA chains impacted by GQDs, respectively [81]. However, these traditional techniques are limited by low observation efficiency and large errors of quantitative results, with are disadvantages in the detection of GFNs [88]. Few studies focus on GFNs nuclear detecting techniques. In the biological imaging field, most research pays attention to safe application of fluorescent GFNs nuclear images rather than assessing genotoxicity of GFNs from an environmental toxicology point of view [91][92][93]. It is necessary to further optimize and develop detection techniques of GFNs in cells and organism tissues for a better understanding of genotoxicity. For example, Chen et al. [94] used laser desorption/ionization mass spectrometry imaging to map and quantify precisely the sub-organ distribution of the carbon nanotubes, GO, and carbon nanodots in mice. The SEM–Raman spectroscopy co-located system provide both SEM and Raman data from the same area on the cell sample, which avoids sample registration issues and makes observed results more accurate [95].

3.2. Genotoxicity Assay of GFNs

There are several assays available to access the genotoxicity of GFNs, measuring various endpoints [96]. The Ames test (bacterial reverse mutation), the comet assay (single cell gel electrophoresis), the chromosomal aberration (CHA), and micronuclei (MN) are the most common tests for genotoxicity. The Ames test (bacterial reverse mutation) can provide initial testing for genotoxicity. The comet assay can detect DNA damage, while the CHA and MN can test large chromosomal abnormalities. The hypoxanthine phosphoribosyl transferase (HPRT) gene is suitable for assessing mutations induced by suspect genotoxic agents, such as NPs [96]. Oxidative DNA damage should be considered one of the causes of genotoxicity. Superoxide radicals can lead to the activation of oxidation of the guanine bases present in the DNA strands, causing rupture to these strands. The most commonly used detection techniques include 8-hydroxydeoxyguanosine and 7, 8-dihydro-oxodeoxyguanine by HPLC with electrochemical detection [97].

References

  1. Chen, T.-T.; Song, W.-L.; Fan, L.-Z. Engineering graphene aerogels with porous carbon of large surface area for flexible all-solid-state supercapacitors. Electrochim. Acta 2015, 165, 92–97.
  2. Kumar, R.; Sahoo, S.; Joanni, E.; Singh, R.K.; Tan, W.K.; Kar, K.K.; Matsuda, A. Recent progress in the synthesis of graphene and derived materials for next generation electrodes of high performance lithium ion batteries. Prog. Energy Combust. Sci. 2019, 75, 100786.
  3. Li, G.; Huang, B.; Pan, Z.; Su, X.; Shao, Z.; An, L. Advances in three-dimensional graphene-based materials: Configurations, preparation and application in secondary metal (Li, Na, K, Mg, Al)-ion batteries. Energy Environ. Sci. 2019, 12, 2030–2053.
  4. Li, Z.; Liu, X.; Wang, L.; Bu, F.; Wei, J.; Pan, D.; Wu, M. Hierarchical 3D all-carbon composite structure modified with N-doped graphene quantum dots for high-performance flexible supercapacitors. Small 2018, 14, 1801498.
  5. Tsai, M.L.; Wei, W.-R.; Tang, L.; Chang, H.-C.; Tai, S.-H.; Yang, P.-K.; Lau, S.P.; Chen, L.-J.; He, J.-H. Si Hybrid Solar Cells with 13% Efficiency via concurrent improvement in optical and electrical properties by employing graphene quantum dots. ACS Nano 2016, 10, 815–821.
  6. Young, C.; Park, T.; Yi, J.W.; Kim, J.; Hossain, M.S.A.; Kaneti, Y.V.; Yamauchi, Y. Advanced functional carbons and their hybrid nanoarchitectures towards supercapacitor applications. Chemsuschem 2018, 11, 3546–3558.
  7. Zhao, Y.-Q.; Zhao, D.-D.; Tang, P.-Y.; Wang, Y.-M.; Xu, C.-L.; Li, H.-L. MnO2/graphene/nickel foam composite as high performance supercapacitor electrode via a facile electrochemical deposition strategy. Mater. Lett. 2012, 76, 127–130.
  8. Zhou, S.; Lin, M.; Zhuang, Z.; Liu, P.; Chen, Z. Biosynthetic graphene enhanced extracellular electron transfer for high performance anode in microbial fuel cell. Chemosphere 2019, 232, 396–402.
  9. Afsahi, S.; Lerner, M.B.; Goldstein, J.M.; Lee, J.; Tang, X.; Bagarozzi, D.A., Jr.; Pan, D.; Locascio, L.; Walker, A.; Barron, F.; et al. Novel graphene-based biosensor for early detection of Zika virus infection. Biosens. Bioelectron. 2018, 100, 85–88.
  10. Chen, H.; Wang, Z.; Zong, S.; Chen, P.; Zhu, D.; Wu, L.; Cui, Y. A graphene quantum dot-based FRET system for nuclear-targeted and real-time monitoring of drug delivery. Nanoscale 2015, 7, 15477–15486.
  11. Iannazzo, D.; Pistone, A.; Ferro, S.; De Luca, L.; Monforte, A.M.; Romeo, R.; Buemi, M.R.; Pannecouque, C. Graphene quantum dots based systems as HIV inhibitors. Bioconjug. Chem. 2018, 29, 3084–3093.
  12. Li, N.; Than, A.; Chen, J.; Xi, F.; Liu, J.; Chen, P. Graphene quantum dots based fluorescence turn-on nanoprobe for highly sensitive and selective imaging of hydrogen sulfide in living cells. Biomater. Sci. 2018, 6, 779–784.
  13. Li, N.; Than, A.; Sun, C.; Tian, J.; Chen, J.; Pu, K.; Dong, X.; Chen, P. Monitoring dynamic cellular redox homeostasis using fluorescence-switchable graphene quantum dots. ACS Nano 2016, 10, 11475–11482.
  14. Li, N.; Than, A.; Wang, X.; Xu, S.; Sun, L.; Duan, H.; Xu, C.; Chen, P. Ultrasensitive profiling of metabolites using tyramine-functionalized graphene quantum dots. ACS Nano 2016, 10, 3622–3629.
  15. Ma, N.; Liu, J.; He, W.; Li, Z.; Luan, Y.; Song, Y.; Garg, S. lFolic acid-grafted bovine serum albumin decorated graphene oxide: An efficient drug carrier for targeted cancer therapy. J. Colloid Interface Sci. 2017, 490, 598–607.
  16. Marzo, A.M.L.; Mayorga-Martinez, C.C.; Pumera, M. 3D-printed graphene direct electron transfer enzyme biosensors. Biosens. Bioelectron. 2020, 151, 111980.
  17. Divyapriya, G.; Vijayakumar, K.K.; Nambi, I. Development of a novel graphene/Co3O4 composite for hybrid capacitive deionization system. Desalination 2019, 451, 102–110.
  18. Gan, L.; Li, H.; Chen, L.; Xu, L.; Liu, J.; Geng, A.; Mei, C.; Shang, S. Graphene oxide incorporated alginate hydrogel beads for the removal of various organic dyes and bisphenol A in water. Colloid Polym. Sci. 2018, 296, 607–615.
  19. Liu, S.; Ge, H.; Wang, C.; Zou, Y.; Liu, J. Agricultural waste/graphene oxide 3D bio-adsorbent for highly efficient removal of methylene blue from water pollution. Sci. Total Environ. 2018, 628–629, 959–968.
  20. Miao, W.; Li, Z.-K.; Yan, X.; Guo, Y.-J.; Lang, W.-Z. Improved ultrafiltration performance and chlorine resistance of PVDF hollow fiber membranes via doping with sulfonated graphene oxide. Chem. Eng. J. 2017, 317, 901–912.
  21. Zhang, D.; Zhao, Y.; Chen, L. Fabrication and characterization of amino-grafted graphene oxide modified ZnO with high photocatalytic activity. Appl. Surf. Sci. 2018, 458, 638–647.
  22. Zhu, Z.; Jiang, J.; Wang, X.; Huo, X.; Xu, Y.; Li, Q.; Wang, L. Improving the hydrophilic and antifouling properties of polyvinylidene fluoride membrane by incorporation of novel nanohybrid 2 particles. Chem. Eng. J. 2017, 314, 266–276.
  23. Afroj, S.; Tan, S.; Abdelkader, A.M.; Novoselov, K.S.; Karim, N. Highly conductive, scalable, and machine washable graphene-based e-textiles for multifunctional wearable electronic applications. Adv. Funct. Mater. 2020, 30, 2000293.
  24. Kurra, N.; Jiang, Q.; Nayak, P.; Alshareef, H.N. Laser-derived graphene: A three-dimensional printed graphene electrode and its emerging applications. Nano Today 2019, 24, 81–102.
  25. Liu, C.; Qiu, S.; Du, P.; Zhao, H.; Wang, L. An ionic liquid-graphene oxide hybrid nanomaterial: Synthesis and anticorrosive applications. Nanoscale 2018, 10, 8115–8124.
  26. Ahmed, F.; Rodrigues, D.F. Investigation of acute effects of graphene oxide on wastewater microbial community: A case study. J. Hazard. Mater. 2013, 256, 33–39.
  27. Arvidsson, R.; Molander, S.; Sanden, B.A. Review of potential environmental and health risks of the nanomaterial graphene. Hum. Ecol. Risk Assess. 2013, 19, 873–887.
  28. Lee, J.H.; Han, J.H.; Kim, J.H.; Kim, B.; Bello, D.; Kim, J.K.; Lee, G.H.; Sohn, E.K.; Lee, K.; Ahn, K.; et al. Exposure monitoring of graphene nanoplatelets manufacturing workplaces. Inhal. Toxicol. 2016, 28, 281–291.
  29. Pelin, M.; Sosa, S.; Prato, M.; Tubaro, A. Occupational exposure to graphene based nanomaterials: Risk assessment. Nanoscale 2018, 10, 15894–15903.
  30. Ou, L.; Song, B.; Liang, H.; Liu, J.; Feng, X.; Deng, B.; Sun, T.; Shao, L. Toxicity of graphene-family nanoparticles: A general review of the origins and mechanisms. Part. Fibre Toxicol. 2016, 13, 57.
  31. Seabra, A.B.; Paula, A.J.; De Lima, R.; Alves, O.L.; Duran, N. Nanotoxicity of graphene and graphene oxide. Chem. Res. Toxicol. 2014, 27, 159–168.
  32. Lee, J.K.; Jeong, A.Y.; Bae, J.; Seok, J.H.; Yang, J.-Y.; Roh, H.S.; Jeong, J.; Han, Y.; Jeong, J.; Cho, W.-S. The role of surface functionalization on the pulmonary inflammogenicity and translocation into mediastinal lymph nodes of graphene nanoplatelets in rats. Arch. Toxicol. 2017, 91, 667–676.
  33. Wu, Y.; Feng, W.; Liu, R.; Xia, T.; Liu, S. Graphene oxide causes disordered zonation due to differential intralobular localization in the liver. ACS Nano 2020, 14, 877–890.
  34. Xie, Y.; Wan, B.; Yang, Y.; Cui, X.; Xin, Y.; Guo, L.-H. Cytotoxicity and autophagy induction by graphene quantum dots with different functional groups. J. Environ. Sci. 2019, 77, 198–209.
  35. Evariste, L.; Lagier, L.; Gonzalez, P.; Mottier, A.; Mouchet, F.; Cadarsi, S.; Lonchambon, P.; Daffe, G.; Chimowa, G.; Sarrieu, C.; et al. Thermal reduction of graphene oxide mitigates its in vivo genotoxicity toward xenopus laevis tadpoles. Nanomaterials 2019, 9, 584.
  36. Feng, X.L.; Zhang, Y.Q.; Luo, R.H.; Lai, X.; Chen, A.J.; Zhang, Y.L.; Hu, C.; Chen, L.L.; Shao, L.Q. Graphene oxide disrupted mitochondrial homeostasis through inducing intracellular redox deviation and autophagy-lysosomal network dysfunction in SH-SY5Y cells. J. Hazard. Mater. 2021, 416, 126158.
  37. Xu, K.; Wang, X.; Lu, C.; Liu, Y.; Zhang, D.; Cheng, J. Toxicity of three carbon-based nanomaterials to earthworms: Effect of morphology on biomarkers, cytotoxicity, and metabolomics. Sci. Total Environ. 2021, 777, 146224.
  38. Barrios, A.C.; Wang, Y.; Gilbertson, L.M.; Perreault, F. Structure-property-toxicity relationships of graphene oxide: Role of surface chemistry on the mechanisms of interaction with bacteria. Environ. Sci. Technol. 2019, 53, 14679–14687.
  39. Li, R.; Guiney, L.M.; Chang, C.H.; Mansukhani, N.D.; Ji, Z.; Wang, X.; Liao, Y.-P.; Jiang, W.; Sun, B.; Hersam, M.C.; et al. Surface oxidation of graphene oxide determines membrane damage, lipid peroxidation, and cytotoxicity in macrophages in a pulmonary toxicity model. ACS Nano 2018, 12, 1390–1402.
  40. Chen, Y.; Ren, C.; Ouyang, S.; Hu, X.; Zhou, Q. Mitigation in multiple effects of graphene oxide toxicity in zebrafish embryogenesis driven by humic acid. Environ. Sci. Technol. 2015, 49, 10147–10154.
  41. Ding, X.; Rui, Q.; Zhao, Y.; Shao, H.; Yin, Y.; Wu, Q.; Wang, D. Toxicity of graphene oxide in nematodes with a deficit in the epidermal barrier caused by RNA interference knockdown of unc-52. Environ. Sci. Technol. Lett. 2018, 5, 622–628.
  42. Liao, K.-H.; Lin, Y.-S.; Macosko, C.W.; Haynes, C.L. Cytotoxicity of graphene oxide and graphene in human erythrocytes and skin fibroblasts. ACS Appl. Mater. Interfaces 2011, 3, 2607–2615.
  43. Bramini, M.; Sacchetti, S.; Armirotti, A.; Rocchi, A.; Vazquez, E.; Leon Castellanos, V.; Bandiera, T.; Cesca, F.; Benfenati, F. Graphene oxide nanosheets disrupt lipid composition, Ca2+ homeostasis, and synaptic transmission in primary cortical neurons. ACS Nano 2016, 10, 7154–7171.
  44. Zhou, J.; Ouedraogo, M.; Qu, F.; Duez, P. Potential genotoxicity of traditional chinese medicinal plants and phytochemicals: An overview. Phytother. Res. 2013, 27, 1745–1755.
  45. Bohne, J.; Cathomen, T. Genotoxicity in gene therapy: An account of vector integration and designer nucleases. Curr. Opin. Mol. Ther. 2008, 10, 214–223.
  46. Huang, R.; Zhou, Y.; Hu, S.; Zhou, P.-K. Targeting and non-targeting effects of nanomaterials on DNA: Challenges and perspectives. Rev. Environ. Sci. Biotechnol. 2019, 18, 617–634.
  47. Burgum, M.J.; Clift, M.J.D.; Evans, S.J.; Hondow, N.; Tarat, A.; Jenkins, G.J.; Doak, S.H. Few-layer graphene induces both primary and secondary genotoxicity in epithelial barrier models in vitro. J. Nanobiotechnol. 2021, 19, 24.
  48. De Souza, T.A.J.; Souza, L.R.R.; Franchi, L.P. Silver nanoparticles: An integrated view of green synthesis methods, transformation in the environment, and toxicity. Ecotoxicol. Environ. Saf. 2019, 171, 691–700.
  49. Mortezaee, K.; Najafi, M.; Samadian, H.; Barabadi, H.; Azarnezhad, A.; Ahmadi, A. Redox interactions and genotoxicity of metal-based nanoparticles: A comprehensive review. Chem. Biol. Interact. 2019, 312, 108814.
  50. Lam, P.L.; Wong, R.S.M.; Lam, K.H.; Hung, L.K.; Wong, M.M.; Yung, L.H.; Ho, Y.W.; Wong, W.Y.; Hau, D.K.P.; Gambari, R.; et al. The role of reactive oxygen species in the biological activity of antimicrobial agents: An updated mini review. Chem. Biol. Interact. 2020, 320, 109023.
  51. Liu, B.; Zhao, Y.; Jia, Y.; Liu, J. Heating Drives DNA to Hydrophobic regions while freezing drives DNA to hydrophilic regions of graphene oxide for highly robust biosensors. J. Am. Chem. Soc. 2020, 142, 14702–14709.
  52. Lan, J.; Gou, N.; Gao, C.; He, M.; Gu, A.Z. Comparative and mechanistic genotoxicity assessment of nanomaterials via a quantitative toxicogenomics approach across multiple species. Environ. Sci. Technol. 2014, 48, 12937–12945.
  53. Cavallo, D.; Fanizza, C.; Ursini, C.L.; Casciardi, S.; Paba, E.; Ciervo, A.; Fresegna, A.M.; Maiello, R.; Marcelloni, A.M.; Buresti, G.; et al. Multi-walled carbon nanotubes induce cytotoxicity and genotoxicity in human lung epithelial cells. J. Appl. Toxicol. 2012, 32, 454–464.
  54. Kim, Y.J.; Rahman, M.M.; Lee, S.M.; Kim, J.M.; Park, K.; Kang, J.-H.; Seo, Y.R. Assessment of in vivo genotoxicity of citrated-coated silver nanoparticles via transcriptomic analysis of rabbit liver tissue. Int. J. Nanomed. 2019, 14, 393–405.
  55. Jarosz, A.; Skoda, M.; Dudek, I.; Szukiewicz, D. Oxidative stress and mitochondrial activation as the main mechanisms underlying graphene toxicity against human cancer cells. Oxid. Med. Cell. Longev. 2016, 2016, 5851035.
  56. Ouyang, S.; Hu, X.; Zhou, Q. Envelopment-internalization synergistic effects and metabolic mechanisms of graphene oxide on single-cell chlorella vulgaris are dependent on the nanomaterial particle size. ACS Appl. Mater. Interfaces 2015, 7, 18104–18112.
  57. Hu, J.; Lin, W.; Lin, B.; Wu, K.; Fan, H.; Yu, Y. Persistent DNA methylation changes in zebrafish following graphene quantum dots exposure in surface chemistry-dependent manner. Ecotoxicol. Environ. Saf. 2019, 169, 370–375.
  58. Zhao, J.; Wang, Z.; White, J.C.; Xing, B. Graphene in the aquatic environment: Adsorption, dispersion, toxicity and transformation. Environ. Sci. Technol. 2014, 48, 9995–10009.
  59. Szabo, P.; Zelko, R. Formulation and stability aspects of nanosized solid drug delivery systems. Curr. Pharm. Des. 2015, 21, 3148–3157.
  60. Singh, N.; Manshian, B.; Jenkins, G.J.S.; Griffiths, S.M.; Williams, P.M.; Maffeis, T.G.G.; Wright, C.J.; Doak, S.H. NanoGenotoxicology: The DNA damaging potential of engineered nanomaterials. Biomaterials 2009, 30, 3891–3914.
  61. Ema, M.; Gamo, M.; Honda, K. A review of toxicity studies on graphene-based nanomaterials in laboratory animals. Regul. Toxicol. Pharmacol. 2017, 85, 7–24.
  62. Ou, L.; Lv, X.; Wu, Z.; Xia, W.; Huang, Y.; Chen, L.; Sun, W.; Qi, Y.; Yang, M.; Qi, L. Oxygen content-related DNA damage of graphene oxide on human retinal pigment epithelium cells. J. Mater. Sci. Mater. Med. 2021, 32, 20.
  63. Poulsen, S.S.; Bengtson, S.; Williams, A.; Jacobsen, N.R.; Troelsen, J.T.; Halappanavar, S.; Vogel, U. A transcriptomic overview of lung and liver changes one day after pulmonary exposure to graphene and graphene oxide. Toxicol. Appl. Pharmacol. 2021, 410, 115343.
  64. Burgum, M.J.; Clift, M.J.D.; Evans, S.J.; Hondow, N.; Miller, M.; Lopez, S.B.; Williams, A.; Tarat, A.; Jenkins, G.J.; Doak, S.H. In vitro primary-indirect genotoxicity in bronchial epithelial cells promoted by industrially relevant few-layer graphene. Small 2021, 17, 2002551.
  65. Hinzmann, M.; Jaworski, S.; Kutwin, M.; Jagiello, J.; Kozinski, R.; Wierzbicki, M.; Grodzik, M.; Lipinska, L.; Sawosz, E.; Chwalibog, A. Nanoparticles containing allotropes of carbon have genotoxic effects on glioblastoma multiforme cells. Int. J. Nanomed. 2014, 9, 2409–2417.
  66. Wu, T.; Liang, X.; Liu, X.; Li, Y.; Wang, Y.; Kong, L.; Tang, M. Induction of ferroptosis in response to graphene quantum dots through mitochondrial oxidative stress in microglia. Part. Fibre Toxicol. 2020, 17, 30.
  67. Wang, A.; Pu, K.; Dong, B.; Liu, Y.; Zhang, L.; Zhang, Z.; Duan, W.; Zhu, Y. Role of surface charge and oxidative stress in cytotoxicity and genotoxicity of graphene oxide towards human lung fibroblast cells. J. Appl. Toxicol. 2013, 33, 1156–1164.
  68. Krasteva, N.; Keremidarska-Markova, M.; Hristova-Panusheva, K.; Andreeva, T.; Speranza, G.; Wang, D.Y.; Draganova-Filipova, M.; Miloshev, G.; Georgieva, M. Aminated graphene oxide as a potential new therapy for colorectal cancer. Oxid. Med. Cell. Longev. 2019, 2019, 3738980.
  69. Hashemi, E.; Akhavan, O.; Shamsara, M.; Rahighi, R.; Esfandiar, A.; Tayefeh, A.R. Cyto and genotoxicities of graphene oxide and reduced graphene oxide sheets on spermatozoa. RSC Adv. 2014, 4, 27213–27223.
  70. Liu, M.; Zhao, H.M.; Chen, S.; Yu, H.T.; Quan, X. Salt-controlled assembly of stacked-graphene for capturing fluorescence and its application in chemical genotoxicity screening. J. Mater. Chem. 2011, 21, 15266–15272.
  71. Chatterjee, N.; Eom, H.-J.; Choi, J. A systems toxicology approach to the surface functionality control of graphene-cell interactions. Biomaterials 2014, 35, 1109–1127.
  72. Sasidharan, A.; Panchakarla, L.S.; Chandran, P.; Menon, D.; Nair, S.; Rao, C.N.R.; Koyakutty, M. Differential nano-bio interactions and toxicity effects of pristine versus functionalized graphene. Nanoscale 2011, 3, 2461–2464.
  73. Franqui, L.S.; De Farias, M.A.; Portugal, R.V.; Costa, C.A.R.; Domingues, R.R.; Souza, A.G.; Coluci, V.R.; Leme, A.F.P.; Martinez, D.S.T. Interaction of graphene oxide with cell culture medium: Evaluating the fetal bovine serum protein corona formation towards in vitro nanotoxicity assessment and nanobiointeractions. Mater. Sci. Eng. C 2019, 100, 363–377.
  74. Borandeh, S.; Alimardani, V.; Abolmaali, S.S.; Seppala, J. Graphene family nanomaterials in ocular applications: Physicochemical properties and toxicity. Chem. Res. Toxicol. 2021, 34, 1386–1402.
  75. Li, J.; Wang, X.; Mei, K.-C.; Chang, C.H.; Jiang, J.; Liu, X.; Liu, Q.; Guiney, L.M.; Hersam, M.C.; Liao, Y.-P.; et al. Lateral size of graphene oxide determines differential cellular uptake and cell death pathways in Kupffer cells, LSECs, and hepatocytes. Nano Today 2021, 37, 101061.
  76. Reina, G.; Ngoc Do Quyen, C.; Nishina, Y.; Bianco, A. Graphene oxide size and oxidation degree govern its supramolecular interactions with siRNA. Nanoscale 2018, 10, 5965–5974.
  77. Syama, S.; Mohanan, P.V. Safety and biocompatibility of graphene: A new generation nanomaterial for biomedical application. Int. J. Biol. Macromol. 2016, 86, 546–555.
  78. Kong, Z.; Hu, W.; Jiao, F.; Zhang, P.; Shen, J.; Cui, B.; Wang, H.; Liang, L. Theoretical evaluation of DNA genotoxicity of graphene quantum dots: A combination of density functional theory and molecular dynamics simulations. J. Phys. Chem. B 2020, 124, 9335–9342.
  79. Li, B.; Zhang, Y.; Meng, X.-Y.; Zhou, R. Zipper-Like unfolding of dsDNA caused by graphene wrinkles. J. Phys. Chem. C 2020, 124, 3332–3340.
  80. Chatterjee, N.; Yang, J.; Choi, J. Differential genotoxic and epigenotoxic effects of graphene family nanomaterials (GFNs) in human bronchial epithelial cells. Mutat. Res. Genet. Toxicol. Environ. Mutagen. 2016, 798, 1–10.
  81. Xu, L.; Dai, Y.; Wang, Z.; Zhao, J.; Li, F.; White, J.C.; Xing, B. Graphene quantum dots in alveolar macrophage: Uptake-exocytosis, accumulation in nuclei, nuclear responses and DNA cleavage. Part. Fibre Toxicol. 2018, 15, 45.
  82. Akhavan, O.; Ghaderi, E.; Akhavan, A. Size-dependent genotoxicity of graphene nanoplatelets in human stem cells. Biomaterials 2012, 33, 8017–8025.
  83. Volkov, Y.; McIntyre, J.; Prina-Mello, A. Graphene toxicity as a double-edged sword of risks and exploitable opportunities: A critical analysis of the most recent trends and developments. 2D Mater. 2017, 4, 022001.
  84. Yao, J.; Wang, H.; Chen, M.; Yang, M. Recent advances in graphene-based nanomaterials: Properties, toxicity and applications in chemistry, biology and medicine. Microchim. Acta 2019, 186, 395.
  85. Akhavan, O.; Ghaderi, E. Toxicity of graphene and graphene oxide nanowalls against bacteria. ACS Nano 2010, 4, 5731–5736.
  86. Domenech, J.; Hernandez, A.; Demir, E.; Marcos, R.; Cortes, C. Interactions of graphene oxide and graphene nanoplatelets with the in vitro Caco-2/HT29 model of intestinal barrier. Sci. Rep. 2020, 10, 2793.
  87. Hu, Q.; Jiao, B.; Shi, X.; Valle, R.P.; Zuo, Y.Y.; Hu, G. Effects of graphene oxide nanosheets on the ultrastructure and biophysical properties of the pulmonary surfactant film. Nanoscale 2015, 7, 18025–18029.
  88. Feng, X.; Chen, Q.; Guo, W.; Zhang, Y.; Hu, C.; Wu, J.; Shao, L. Toxicology data of graphene-family nanomaterials: An update. Arch. Toxicol. 2020, 94, 1915–1939.
  89. Roth, G.A.; Tahiliani, S.; Neu-Baker, N.M.; Brenner, S.A. Hyperspectral microscopy as an analytical tool for nanomaterials. Wiley Interdiscip. Rev. Nanomed. Nanobiotechnol. 2015, 7, 565–579.
  90. Ouyang, S.; Zhou, Q.; Zeng, H.; Wang, Y.; Hu, X. Natural nanocolloids mediate the phytotoxicity of graphene oxide. Environ. Sci. Technol. 2020, 54, 4865–4875.
  91. Schroeder, K.L.; Goreham, R.V.; Nann, T. Graphene quantum dots for theranostics and bioimaging. Pharm. Res. 2016, 33, 2337–2357.
  92. Wu, C.; Wang, C.; Han, T.; Zhou, X.; Guo, S.; Zhang, J. Insight into the cellular internalization and cytotoxicity of graphene quantum dots. Adv. Healthc. Mater. 2013, 2, 1613–1619.
  93. Gao, W.; Zhang, J.; Xue, Q.; Yin, X.; Yin, X.; Li, C.; Wang, Y. Acute and subacute toxicity study of graphene-based tumor cell nucleus-targeting fluorescent nanoprobes. Mol. Pharm. 2020, 17, 2682–2690.
  94. Chen, S.; Xiong, C.; Liu, H.; Wan, Q.; Hou, J.; He, Q.; Badu-Tawiah, A.; Nie, Z. Mass spectrometry imaging reveals the sub-organ distribution of carbon nanomaterials. Nat. Nanotechnol. 2015, 10, 176–182.
  95. Cardell, C.; Guerra, I. An overview of emerging hyphenated SEM-EDX and Raman spectroscopy systems: Applications in life, environmental and materials sciences. TrAC Trends Anal. Chem. 2016, 77, 156–166.
  96. Magdolenova, Z.; Collins, A.; Kumar, A.; Dhawan, A.; Stone, V.; Dusinska, M. Mechanisms of genotoxicity. A review of in vitro and in vivo studies with engineered nanoparticles. Nanotoxicology 2014, 8, 233–278.
  97. Ganguly, P.; Breen, A.; Pillai, S.C. Toxicity of nanomaterials: Exposure, pathways, assessment, and recent advances. ACS Biomater. Sci. Eng. 2018, 4, 2237–2275.
More
Information
Contributor MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register :
View Times: 326
Revisions: 2 times (View History)
Update Date: 10 Nov 2021
1000/1000
Video Production Service