015731320263Acc Chem ResAcc. Chem. Res.Accounts of chemical research0001-48421520-489822824066347315810.1021/ar300106hNIHMS396149ArticleA Natural Vanishing Act: The Enzyme-Catalyzed Degradation of Carbon NanomaterialsKotcheyGregg P.1HasanSaad A.1KapralovAlexander A.2HaSeung Han3KimKang3ShvedovaAnna A.45KaganValerian E.2StarAlexander1*Department of Chemistry, University of Pittsburgh, Pittsburgh, PennsylvaniaDepartment of Environmental and Occupational Health, University of Pittsburgh, Pittsburgh, PennsylvaniaDepartment of Medicine, University of Pittsburgh, Pittsburgh, PennsylvaniaPathology and Physiology Research Branch, Health Effects Laboratory Division, National Institute for Occupational Safety and Health (NIOSH/CDC), Morgantown, West VirginiaDepartment of Physiology and Pharmacology, School of Medicine, West Virginia Un iversity, Morgantown, WVTo whom all correspondence should be addressed. astar@pitt.edu247201223720121610201216102013451017701781CONSPECTUS

Over the past three decades, revolutionary research in nanotechnology by the scientific, medical, and engineering communities has yielded a treasure trove of discoveries with diverse applications that promise to benefit humanity. With their unique electronic and mechanical properties, carbon nanomaterials (CNMs) represent a prime example of the promise of nanotechnology with applications in areas that include electronics, fuel cells, composites, and nanomedicine. Because of toxicological issues associated with CNMs, however, their full commercial potential may not be achieved. The ex vitro, in vitro, and in vivo data presented in this Account provide fundamental insights into the biopersistence of CNMs, such as carbon nanotubes and graphene, and their oxidation/biodegradation processes as catalyzed by peroxidase enzymes. We also communicate our current understanding of the mechanism for the enzymatic oxidation/biodegradation. Finally, we outline potential future directions that could enhance our mechanistic understanding of the CNM oxidation/biodegradation and could yield benefits in terms of human health and environmental safety.

The conclusions presented in this Account may catalyze a rational rethinking of CNM incorporation in diverse applications. For example, armed with an understanding of how and why CNMs undergo enzyme-catalyzed oxidation/biodegradation, researchers can tailor the structure of CNMs to either promote or inhibit these processes. In nanomedical applications such as drug delivery, the incorporation of carboxylate functional groups could facilitate biodegradation of the nanomaterial after delivery of the cargo. On the other hand, in the construction of aircraft, a CNM composite material should be stable to oxidizing conditions in the environment. Therefore, pristine, inert CNMs would be ideal for this application. Finally, the incorporation of CNMs with defect sites in consumer goods could provide a facile mechanism that promotes the degradation of these materials once these products reach landfills.

Biodegradationenzymeoxidationcarbon nanotubesgraphenehorseradish peroxidasemyeloperoxidaseNational Institute of Allergy and Infectious Diseases Extramural Activities : NIAIDU19 AI068021 || AINational Institute of Allergy and Infectious Diseases Extramural Activities : NIAIDU19 AI068021 || AINational Heart, Lung, and Blood Institute : NHLBIR01 HL094488 || HLNational Heart, Lung, and Blood Institute : NHLBIR01 HL070755 || HLNational Institute of Environmental Health Sciences : NIEHSR01 ES019304 || ES
Introduction

Carbon nanotubes and graphene, the nanoscale sp2 allotropes of carbon, have garnered widespread attention as a result of their remarkable electrical, mechanical, and optical properties and the promise of new technologies that harness these properties. Today, these breakthroughs no longer remain just a vision. The Project on Emerging Nanotechnologies has documented over 30 products containing these carbon nanomaterials (CNMs), in the form of composites, dispersions, and solid-state devices, from a broad selection of industries: aerospace, consumer electronics, cosmetics, and sporting goods.1 The manufacture and eventual disposal of these products may result in the release of CNMs into the environment and subsequent exposure of humans, animals, and vegetation to these CNMs. Given possible pro-inflammatory and toxic effects of CNMs, much attention has been focused on their health effects and interactions with living systems,26 with other review articles addressing the distribution, toxicity, and persistence of CNMs in the environment.79

Recently, several groups have reported the degradation of CNMs catalyzed by naturally occurring enzymes from plants and humans. These findings suggest that the potentially detrimental effects of CNMs associated with their biopersistence may be mitigated thereby enabling these materials to be used more widely in a safe manner. Here, we review these findings and discuss possible mechanisms of enzymatic degradation. We relate the oxidative potencies of these enzymes to those of inorganic oxidants employed to oxidize/degrade CNMs. We further elucidate the role of chemical functionalization in CNM degradation. Finally, we discuss data relating to the oxidative biodegradation of CNMs both in vitro10 and in vivo11 as potential means to mitigate the distribution and toxicological effects of CNMs after exposure.

Chemical Oxidation of Carbon Nanomaterials

The scientific literature is rich with reports regarding the covalent modification of CNMs1215 with those relating to oxidizing processes summarized in Scheme 1. Strong oxidants, such as H2SO4 and HNO3, disrupt the sp2 carbon lattice of unfunctionalized CNMs and introduce oxygen-containing functional groups. Typical morphological changes accompanying such oxidation are a decrease in length for nanotubes16 and the appearance of holes in graphene sheets.17 Sustained oxidation leads to a loss of material from the carbon framework in the form of by-products such as CO, CO2, as well as aliphatic and polyaromatic hydrocarbons.10,18 While it was clearly demonstrated in the literature that strong oxidants can oxidize CNMs, these nanomaterials were largely considered to be biopersistent under physiological conditions and in the environment. We, however, hypothesized that like biodegradable polymers CNMs may undergo enzymatic oxidative modifications and biodegradation both in vivo and in the environment.

Introduction to Peroxidase Enzymes Utilized for CNM Degradation/Biodegradation

To date, primarily two peroxidase enzymes, horseradish peroxidase (HRP, Figure 1A) and myeloperoxidase (MPO, Figure 1B) have been studied for the degradation/biodegradation of CNMs. For over a century, HRP (molecular weight ~44 kDa), an enzyme secreted by the root of the horseradish plant (Armoracia rusticana), has been utilized in chemical synthesis, biotechnology, and bioremediation.19 Another enzyme, MPO, consisting of two identical dimers with a total weight ~144 kDa, is expressed in inflammatory cells (predominantly neutrophils) of animals and humans. MPO generates reactive intermediates and oxidants that serve to destroy infecting bacteria.2022

Both HRP and MPO contain a heme group as their active sites, which is a Fe(III) protoporphyrin IX (Figure 1C). When the enzymes are inactive, the heme group is in the ferric resting state (Figure 1D). In the presence of hydrogen peroxide (H2O2), the heme group undergoes a protein-assisted conversion to a ferryl oxo iron (Fe4+=O) porphyrin p cation radical known as Compound I, which is subsequently returned to the ferric resting state in two sequential, one-electron transfer steps.19,22 In the first step, the transient intermediate, Compound II, forms when the porphyrin p cation radical is reduced as a substrate (AH2) is oxidized.19,22 AH2 is further oxidized when the ferryl oxo iron is reduced to the ferric resting state.19,22 Under the acidic conditions of the neutrophil and in the presence of chloride (Cl), Compound I of MPO abstracts two electrons from Cl to yield hypochlorous acid (HOCl) as it returns to the ferric resting state.22

Table 1 lists the redox potentials of HRP and MPO. While HRP is a weaker oxidant than MPO, the redox potentials of HRP and MPO are comparable to those of nitric acid (0.957 V) and ozone (1.24 V), respectively.23 Moreover, HOCl, which has a redox potential of 1.63 V,23 functions as a stronger oxidant than MPO at each of its steps of the peroxidase cycle.

HRP/H<sub>2</sub>O<sub>2</sub>-Catalyzed Degradation of CNMs

We demonstrated that carboxylated single-walled carbon nanotubes (SWCNTs), underwent morphological changes (shortening and deformation) when statically incubated at 4°C in the presence of HRP and H2O2 (~40 μM).25 Utilizing transmission electron microscopy (TEM), SWCNTs with a mean length ~500 nm (Figure 2A) were observed to shorten by ~45% over 8 weeks, and mostly residual globular material was detected at 12 weeks (Figure 2B). In contrast, pristine SWCNTs failed to degrade when incubated with HRP and H2O2 (~80 μM), from which we theorized that defects or functionalized sites were important facilitators of enzymatic action.26 Bianco and coworkers corroborated these findings in a study of HRP/H2O2 treatment over 30 days, using Raman spectroscopy to monitor changes in the carbon lattice (Figure 2C).27 The initial Raman spectrum exhibited a strong D band, which was attributed to SWCNT carboxylation. After 7 days of incubation, the D band weakened, suggesting that highly carboxylated nanotubes were oxidized first; on day 30, no Raman signal was observed because the nanotubes were completely oxidized. During the degradation process, the SWCNTs were observed to open via an unzipping mechanism yielding graphene sheets. This result was consistent with recent reports by Tour and coworkers28 and Dai and coworkers,29 which revealed that chemical treatments could longitudinally unzip both SWCNTs and MWCNTs to produce graphene nanoribbons (GNRs, Scheme 2).

More recently, the investigation of CNM degradation was expanded to multiwalled carbon nanotubes (MWCNTs). In one study, Bianco and coworkers subjected chemically carboxylated MWCNTs to HRP/H2O2 treatment, and over 60 days, MWCNTs were observed to shorten between 25 to 63% (Figure 3A). The observed shortening contrasts the unzipping observed for SWCNTs treated under identical conditions.27 Independently, during an 80 day study, we investigated the process of MWCNT degradation catalyzed by HRP/H2O2.30 TEM, including high-resolution imaging, revealed that carboxylated MWCNTs decreased in both length and diameter (Figure 3B), but unlike SWCNTs treated with HRP/H2O2, MWCNTs failed to disappear.30 The decrease in length, in agreement with the report by Bianco and coworkers,27 suggested that degradation proceeded from the ends of the nanotubes. The concurrent decrease in diameter could be explained by exfoliation of defect-containing outer walls, which were unzipped via the enzymatic oxidation, yielding nanotubes with smaller diameters and possessing mostly pristine outer walls (Figure 3C). The exposure of these newly revealed, more pristine layers coincided with a slowdown in the rate of degradation. Again, we could infer that defects or functionalized sites were important facilitators of enzymatic action. This notion was further supported by our experiments with nitrogen-doped MWCNTs, which were effectively degraded by the HRP/H2O2 system over 90 days.30

The importance of functionalization for CNM oxidation/degradation by HRP/H2O2 was further supported by experiments with graphene oxide (GO), a two-dimensional sheet of oxidized graphitic carbon comprising one layer of carbon atoms decorated with oxygen functionalities like epoxides and carboxylates (Figure 4A). GO incubated with HRP/H2O2 was observed via TEM and Raman spectroscopy to undergo enzymatic oxidation.31 After 8 days of incubation, with daily additions of H2O2, holes with a diameter of 2.1 ± 0.6 nm formed on the basal plane of the GO sheet.31 By day 10, the diameter of the holes increased more than 12-fold to 26.7 ± 12.8 nm (Figure 4B), and by day 20, a majority of the GO was oxidized completely. Raman spectroscopy corroborated the TEM observations. In the first 4 days, the D/G intensity ratio increased from 1.1 to 1.6, and on day 20, both the D and G bands disappeared from the signal, indicating progressive oxidation leading to the elimination of the graphitic lattice (Figure 4C). For comparison, the effects of HRP/H2O2 treatment on a chemically reduced graphene oxide (RGO) were studied. RGO has the same single-sheet geometry of GO but is mostly devoid of GO’s oxygen-containing functional groups. In contrast to GO, the HRP/H2O2 treatment failed to oxidize RGO (Figure 4D). Thus, the defect sites introduced by chemical oxidation appeared to have a significant role in the subsequent enzymatic oxidation of the CNM, irrespective of its geometry.

MPO-Catalyzed Degradation of CNMs

Having examined CNM degradation catalyzed by HRP/H2O2, we turn to the report of MPO-catalyzed degradation. With the knowledge that human myeloperoxidase (hMPO) degrades implantable polymers such as poly(ester-urea-urethane), we hypothesized that hypochlorite and reactive radical intermediates generated by hMPO would catalyze the oxidation of carboxylated SWCNTs. Nanotube suspensions incubated with the oxidants hMPO/H2O2 or hypochlorite appeared lighter in color after 24 hours, indicating a decrease in graphitic material (Figure 5A). Moreover, the SWCNT suspension in which NaCl was added to hMPO/H2O2 exhibited significant oxidation as evidenced by its nearly clear appearance after 24 hours. This result suggested an important role of the strong oxidant, hypochlorous acid, in MPO-catalyzed degradation of CNMs.32 Raman spectroscopy conducted at intervals up to 24 h showed a progressive decrease in the G band until it was quenched, indicating complete degradation of graphitic lattice (Figure 5B). Vis–NIR absorbance spectroscopy revealed a suppression of the characteristic semiconducting (S2) and metallic (M1) transition bands of SWCNTs in the same interval (Figure 5C). Finally, scanning electron microscopy (SEM) provided direct observation of the morphological changes to the SWCNTs resulting from the different incubation conditions (Figure 5D). For example, after the hMPO/H2O2/NaCl treatment, the cylindrical structure that characterizes SWCNTs disappeared (Figure 5D, center left); small quantities of carbonaceous residue remained after incubation with hMPO/H2O2 (Figure 5D, center right). Together these measurements showed that hMPO with H2O2 catalyzed the degradation of carboxylated SWCNTs, a process further accelerated by the formation of hypochlorite.

The Role of Functional Groups in Enzymatic Oxidation/Degradation

Based on the literature, Table 2 summarizes the reported degradation of CNMs catalyzed by different enzymes. The different functional groups on the CNMs are noted. In a later section, in vitro and in vivo biodegradation of CNMs will be explored.

A few trends can be gleaned from the ex vitro data presented in Table 2. First, minimal enzyme-catalyzed oxidation of pristine (i.e. nonfunctionalized) CNTs and graphene (i.e. RGO) was demonstrated by the available characterization techniques. Alternatively, carboxylated CNTs incubated with either HRP or MPO were further oxidized. Kane and coworkers proposed a mechanism that correlated CNT oxidation to the disruption of the graphitic lattice incurred by functionalization.34 Strong oxidizing acids employed to functionalize SWCNTs typically yield epoxide, peroxide, hydroxyl, carbonyl, and carboxylate functional groups.34 During carboxylation, three bonds belonging to the participating carbon atoms are broken, rendering this functionality attached to the graphitic backbone by only one bond (Scheme 3). Thus, carboxylated CNTs appeared most susceptible to enzymatic oxidation, because only one bond needed to be cleaved. This assertion was further supported by the results shown in Figure 3 for MWCNTs and their proposed oxidation mechanism. After the outer layers containing functional groups (i.e. carboxylates) were enzymatically oxidized and exfoliated, the more pristine inner layers failed to undergo oxidation.

Enzyme-Substrate Interactions in CNM Oxidation/Degradation

We hypothesized that the carboxylates are important for one additional reason: their negative charge in aqueous media facilitates the proper orientation/proximity of the peroxidase enzyme’s heme active site to the SWCNT substrate. We employed molecular modeling studies to characterize the role of functional groups in enzyme-substrate interactions. A calculation of the docking conformations for carboxylated SWCNTs on HRP predicted that the carboxylated ends, which are negatively charged in an aqueous suspension, would orient themselves toward a positively charged arginine residue, Arg178, near HRP’s heme site (Figure 6A).26 In contrast, pristine SWCNTs, which failed undergo degradation, were predicted to orient themselves with the distal end of the enzyme, away from the active heme site. Similarly, for MPO, molecular simulations suggested two possible sites for interaction between the enzyme and SWCNTs.10 The first site, located at the proximal end of the heme group, contained catalytically active tyrosine residues, Tyr 293 and 313, and arginine residues, Arg 294, 307, and 507;10 again, the positively charged arginine residues interact strongly with the negatively charged carboxylated nanotubes (Figure 6B). The second site, at the distal end of the heme group and distant from the tyrosine residues, appeared to bind the pristine SWCNTs.

The molecular modeling of HRP was expanded from the docking of one-dimensional SWCNTs to that of two-dimensional graphene sheets.31 For this geometry, favorable binding poses were identified only on graphene’s basal plane (Figure 6C) because the edges of the sheet failed to provide sufficient surface area for graphene-HRP stabilization through hydrophobic interactions. Along the basal plane, the heme active site was situated approximately 8.7 Å from the surface of GO. In contrast, the separation distance for RGO was 11.5 Å. The presence of epoxide and alcohol groups on the basal plane of GO may be responsible for orienting the heme active site with enough proximity to the GO thereby enabling its oxidation catalyzed by HRP/H2O2.

Mechanism of the Enzyme-catalyzed CNM Oxidation/Degradation

Scheme 4 depicts the work function under vacuum (V, ~5 eV) and standard potential (E0, +0.5 V) of the valence band for SWCNTs.35,36 Since HRP, MPO, and MPO’s byproduct, HOCl, are stronger oxidants (i.e. demonstrate more positive redox potentials), these species should be spontaneously reduced under standard conditions by accepting electrons from the valence band of SWCNTs and thereby oxidizing them. While thermodynamically HRP and MPO have the potential to spontaneously oxidize SWCNTs, in reality CNT oxidation/degradation may be influenced by enzyme/substrate interactions and is facilitated by defects.

It has been demonstrated that hypochlorite results in the introduction of acid “defects” sites on the surface of CNTs.37 Therefore, for MPO-catalyzed biodegradation, it is possible that the initial oxidation of CNMs is predominantly due to the effects of hypochlorite (as a result of its very high oxidizing potential); the accumulation of oxidative “defects” may pave the way for subsequent more effective oxidation via both mechanisms (i.e. hypochlorite and peroxidase reactive intermediates with lower oxidizing potentials).

What is evident from these proposed mechanisms is the need for further investigations to elucidate the enzyme-catalyzed degradation process at the molecular level. By identifying the by-products of degradation, it will be possible to construct more specific reaction schemes describing the mechanism of the enzyme-catalyzed processes. While this Account has focused on the heterolytic bond cleavage of H2O2 catalyzed by peroxidase enzymes, an alternative mechanism involving the homolytic cleavage of H2O2 catalyzed by transition metals such as iron can also impact CNMs. For example, Smalley and coworkers demonstrated that a metal-catalyzed process could be employed for the purification of SWCNTs,38 and we established that the radicals generated through the homolytic cleavage of H2O2 in the presence of iron and hemin (i.e. via the Fenton reaction) can degrade SWCNTs.26

Biological Investigations – <italic>in vitro</italic> and <italic>in vivo</italic> Biodegradation of CNMs

While ex vitro experiments with HRP and MPO demonstrated the ability of these enzymes to catalyze degradation of CNMs, they did not address the significance of these reactions in vitro and in vivo. Thus, the enzymatic oxidation of CNMs has been studied in living systems. For carbon nanotubes (CNTs), we employed Raman spectroscopy to monitor carboxylated SWCNTs located within neutrophils, inflammatory cells containing high concentrations of MPO. The intensity of the D band relative to the G band increased noticeably going from 2 h to 8 h of incubation within the neutrophils (Figure 7A, B). This change suggests the introduction of defects to the nanotubes via enzymatic oxidation. Neves et al. tracked the process further by measuring the Raman signal from oxidized, RNA-wrapped double-walled CNTs incubated with human prostate adenocarcinoma cells for up to 18 h.39 They reported a steady increase in the D/G intensity ratio with incubation time (Figure 7C) and inferred that this progression was due to defect-inducing alterations in the outer wall of the nanotubes. Unfortunately, they did not pinpoint the actor or mechanism within the cells that altered the structure of the nanotubes’ wall. These are interesting experiments in that RNA is also negatively charged; therefore, the positively charged domains in the putative catalyst would act similarly to the peroxidases described above.

Our in vivo murine experiments afforded the most important evidence for the role of MPO in oxidative biodegradation of CNTs. In this study, the pulmonary inflammation response of mice exposed to carboxylated SWCNTs via pharyngeal aspiration was compared between wild-type (w/t) mice and MPO-knockout (k/o) mice (i.e. mice with and without MPO, respectively).11 One of the primary challenges in assessing CNT biodegradation entails quantitatively evaluating the content of the CNM in vivo. An interesting opportunity is offered by the Hyperspectral Imaging methodology (CytoViva) whereby spectral signatures characteristic of CNTs are collected into a spectral library and subsequently utilized for the detection of CNT in tissue sections (Figure 8A). Furthermore, photoacoustic (PA) imaging technology40 was utilized to assess the content of SWCNTs in sections of the lungs from w/t and MPO-k/o mice (Figure 8B). The decreased PA intensity 28 days after pharyngeal aspiration of SWCNTs was statistically significant in w/t mice but not in MPO-k/o mice (Figure 8C). Combined with the data obtained by several additional quantitative protocols (Raman microscopy, NIR spectroscopy, TEM analysis), we concluded the inflammatory response in MPO k/o mice was stronger than in w/t C57B1/6 mice, and that SWCNT oxidation and clearance from the lungs of MPO k/o animals was markedly less effective (Figure 8D).11 Interestingly, a recent report documented in vivo degradation of CNT in the brain of animals.41

Outlook and Conclusions

Biopersistence of CNMs has been long viewed as the major factor contributing to the toxic effects of these nanomaterials in the body. The discovery of the enzymatic CNT degradation processes,10,11,2527,3032 opened new opportunities for the regulation of CNT distribution and fate in vivo by controlling inflammatory response and/or employing SWCNT-metabolizing enzymes. Indeed, the number of studies relating to the field of CNMs and biodegradation has grown vigorously. For example, PubMed revealed 16 entries using the search terms “biodegrade/biodegradation” and “carbon nanotubes” before 2008; an additional 48 publications have been added since then. Although research in the field has greatly expanded, significant amount of work is still warranted. To date, most research on enzyme-catalyzed biodegradation has focused on HRP ex vitro and MPO in vitro and in vivo; it is possible that other enzymes such as eosinophil peroxidases (EPO) and lactoperoxidase (LPO) could be involved in the biodegradation process. Furthermore, other hemo-proteins with peroxidase activity merit investigation, including different forms of cytochrome P450, cytochrome c, etc. The specific roles and significance of these enzymes in biodegradation and regulation of inflammatory responses should be explored. Additionally, the assessment of the intermediates of biodegradation represents another area of research that merits investigation because the likely products—aliphatic and oxidized polyaromatic hydrocarbons—will provide details about the mechanism of CNM biodegradation and may exert their own specific, possibly toxic, effects on the body, which may also have to be taken into account.

As a result of the revolutionary research on enzyme-catalyzed biodegradation, one can envision futuristic applications in areas like drug delivery or imaging applications. For example, enzymes that result in biodegradation could be encapsulated inside nitrogen-doped nanocapsules,42 a CNM that could be employed to deliver drug or imaging cargo along with all the components for its self-elimination from the body. More practically, relatively non-toxic and inexpensive enzymes may be instrumental in environmental/occupational biodegradation of CNMs contamination introduced via spills, waste products in landfills, etc. To this end, white-rot basidiomycete fungi that secrete lignin peroxidase (LiP) were exposed to a dispersion of C60(OH)n and proceeded to oxidize this CNM over 32 weeks.43 This result suggests the existence of a bioremediation pathway for oxidized CNMs that might otherwise accumulate in the environment. Finally, an interesting avenue of research will be applying enzymatic biodegradation to other non-carbonaceous nanomaterials to mitigate toxicity.

This work was supported by NIEHS R01ES019304, HL70755, HL094488, U19AI068021, NIOSH OH008282 and NORA/NTRC NIOSH 1927ZHF. GPK acknowledges support from the EPA STAR Graduate Fellowship FP-91713801.

Disclaimer. The findings and conclusions in this report are those of the authors and do not necessarily represent the views of the National Institute for Occupational Safety and Health.

The project on emerging nanotechnologieshttp://www.nanotechproject.org/accessed April 9, 2012KaganVETyurinaYYTyurinVAKonduruNVPotapovichAIOsipovANKisinERSchwegler-BerryDMercerRCastranovaVShvedovaAADirect and indirect effects of single walled carbon nanotubes on raw 264.7 macrophages: Role of ironToxicol Lett20061658810016527436TkachAVShurinGVShurinMRKisinERMurrayARYoungSHStarAFadeelBKaganVEShvedovaAADirect effects of carbon nanotubes on dendritic cells induce immune suppression upon pulmonary exposureACS Nano201155755576221657201ShvedovaAAKisinERMercerRMurrayARJohnsonVJPotapovichAITyurinaYYGorelikOArepalliSSchwegler-BerryDHubbsAFAntoniniJEvansDEKuBKRamseyDMaynardAKaganVECastranovaVBaronPUnusual inflammatory and fibrogenic pulmonary responses to single-walled carbon nanotubes in miceAm J Physiol2005289L698L708KisinERMurrayARKeaneMJShiXCSchwegler-BerryDGorelikOArepalliSCastranovaVWallaceWEKaganVEShvedovaAASingle-walled carbon nanotubes: Geno- and cytotoxic effects in lung fibroblast v79 cellsJ Toxicol Environ Health, Part A2007702071207918049996ShvedovaAAFabisiakJPKisinERMurrayARRobertsJRTyurinaYYAntoniniJMFengWHKommineniCReynoldsJBarchowskyACastranovaVKaganVESequential exposure to carbon nanotubes and bacteria enhances pulmonary inflammation and infectivityAm J Respir Cell Mol Biol20083857959018096873PérezSFarréMLBarcelóDAnalysis, behavior and ecotoxicity of carbon-based nanomaterials in the aquatic environmentTrAC-Trend Anal Chem200928820832PetersenEJZhangLMattisonNTO’carrollDMWheltonAJUddinNNguyenTHuangQHenryTBHolbrookRDChenKLPotential release pathways, environmental fate, and ecological risks of carbon nanotubesEnviron Sci Technol2011459837985621988187EckelmanMJMauterMSIsaacsJElimelechMNew perpsectives on nanomaterial aquatic ecotoxicity: Production impacts exceed direct exposure impacts for carbon nanotubesEnviron Sci Technol2012462902291022296240KaganVEKonduruNVFengWAllenBLConroyJVolkovYVlasovaIIBelikovaNAYanamalaNKapralovATyurinaYYShiJKisinERMurrayARFranksJStolzDGouPKlein-SeetharamanJFadeelBStarAShvedovaAACarbon nanotubes degraded by neutrophil myeloperoxidase induce less pulmonary inflammationNat Nanotechnol2010535435920364135ShvedovaAAKapralovAAFengWHKisinERMurrayARMercerRRSt CroixCMLangMAWatkinsSCKonduruNVAllenBLConroyJKotcheyGPMohamedBMMeadeADVolkovYStarAFadeelBKaganVEImpaired clearance and enhanced pulmonary inflammatory/fibrotic response to carbon nanotubes in myeloperoxidase-deficient micePLoS One20127e3092322479306PengXWongSSFunctional covalent chemistry of carbon nanotube surfacesAdv Mater200921625642KarousisNTagmatarchisNTasisDCurrent progress on the chemical modification of carbon nanotubesChem Rev20101105366539720545303LohKPBaoQAngPKYangJThe chemistry of grapheneJ Mater Chem20102022772289DreyerDRParkSBielawskiCWRuoffRSThe chemistry of graphene oxideChem Soc Rev20103922824020023850LiuJRinzlerAGDaiHJHafnerJHBradleyRKBoulPJLuAIversonTShelimovKHuffmanCBRodriguez-MaciasFShonYSLeeTRColbertDTSmalleyREFullerene pipesScience1998280125312569596576ZhaoXHaynerCMKungMCKungHHFlexible holey graphene paper electrodes with enhanced rate capability for energy storage applicationsACS Nano201158739874921980979WangZShirleyMDMeikleSTWhitbyRLDMikhalovskySVThe surface acidity of acid oxidised multi-walled carbon nanotubes and the influence of in-situ generated fulvic acids on their stability in aqueous dispersionsCarbon2009477379VeitchNCHorseradish peroxidase: A modern view of a classic enzymePhytochemistry20046524925914751298NauseefWMHow human neutrophils kill and degrade microbes: An integrated viewImmunol Rev20072198810217850484HanssonMOlssonINauseefWMBiosynthesis, processing, and sorting of human myeloperoxidaseArch Biochem Biophys200644521422416183032ArnholdJProperties, functions, and secretion of human myeloperoxidaseBiochemistry (Mosc)2004694914972011LideDRPotentialsCRC PressBoca Raton20042029HayashiYYamazakiIThe oxidation-reduction potentials of compound i/compound ii and compound ii/ferric couples of horseradish peroxidases a2 and cJ Biol Chem19792549101910639073AllenBLKichambarePDGouPVlasovaIiKapralovAAKonduruNKaganVEStarABiodegradation of single-walled carbon nanotubes through enzymatic catalysisNano Lett200883899390318954125AllenBLKotcheyGPChenYYanamalaNVKKlein-SeetharamanJKaganVEStarAMechanistic investigations of horseradish peroxidase-catalyzed degradation of single-walled carbon nanotubesJ Am Chem Soc2009131171941720519891488RussierJMenard-MoyonCVenturelliEGravelEMarcolongoGMeneghettiMDorisEBiancoAOxidative biodegradation of single- and multi-walled carbon nanotubesNanoscale2011389389621116547KosynkinDVHigginbothamALSinitskiiALomedaJRDimievAPriceBKTourJMLongitudinal unzipping of carbon nanotubes to form graphene nanoribbonsNature200945887287619370030JiaoLZhangLWangXDiankovGDaiHNarrow graphene nanoribbons from carbon nanotubesNature200945887788019370031ZhaoYAllenBLStarAEnzymatic degradation of multiwalled carbon nanotubesJ Phys Chem A20111159536954421348486KotcheyGPAllenBLVedalaHYanamalaNKapralovAATyurinaYYKlein-SeetharamanJKaganVEStarAThe enzymatic oxidation of graphene oxideACS Nano201152098210821344859VlasovaISokolovAChekanovAKostevichVVasilyevVMyeloperoxidase-induced biodegradation of single-walled carbon nanotubes is mediated by hypochloriteRuss J Bioorg Chem201137453463BiancoAKostarelosKPratoMMaking carbon nanotubes biocompatible and biodegradableChem Commun2011471018210188LiuXHurtRHKaneABBiodurability of single-walled carbon nanotubes depends on surface functionalizationCarbon2010481961196920352066GratzelMPhotoelectrochemical cellsNature200141433834411713540ChoiHCShimMBangsaruntipSDaiHSpontaneous reduction of metal ions on the sidewalls of carbon nanotubesJ Am Chem Soc20021249058905912149003YangJCYenCHWangWJHorngJJTsaiYPAssessment of adequate sodium hypochlorite concentration for pre-oxidization of multi-walled carbon nanotubesJ Chem Technol Biot201085699707WangShanHHaugeRHPasqualiMSmalleyREA highly selective, one-pot purification method for single-walled carbon nanotubesJ Phys Chem B20071111249125217249726NevesVHeisterECostaSTîlmaciuCBorowiak-PalenEGiuscaCEFlahautESoulaBColeyHMMcfaddenJSilvaSRPUptake and release of double-walled carbon nanotubes by mammalian cellsAdv Funct Mater20102032723279KimKHuangSWAshkenaziSO’donnellMAgarwalAKotovNADennyMFKaplanMJPhotoacoustic imaging of early inflammatory response using gold nanorodsAppl Phys Lett200790223901223903NunesABussyCGherardiniLMeneghettiMHerreroMABiancoAPratoMPizzorussoTAl-JamalKTKostarelosKIn vivo degradation of functionalized carbon nanotubes after stereotactic administration in the brain cortexNanomedicine201210.2217/nnm.12.33AllenBLShadeCMYinglingAMPetoudSStarAGraphitic nanocapsulesAdv Mater20092146924695SchreinerKMFilleyTRBlanchetteRABowenBBBolskarRDHockadayWCMasielloCARaebigerJWWhite-rot basidiomycete-mediated decomposition of C60 fullerolEnviron Sci Technol2009433162316819534129

Gregg P. Kotchey received his B.S. degree in chemistry (2004) from the University of Pittsburgh, where he is now working on his Ph.D. under the direction of Professor Alexander Star.

Saad A. Hasan received his Ph.D. in materials science (2010) from Vanderbilt University, after which he joined the Nanommune project at Uppsala University (Sweden). Saad was a postdoctoral associate in Professor Star’s group at Pittsburgh.

Alexander A. Kapralov received his Ph.D. in Biochemistry (1986) from Institute of Biochemistry Ukrainian Academy of Sciences (Ukraine) and Dr. Sc. in Biochemistry (2000) from National Kyiv University (Ukraine). He is now a research associate with Professor Kagan at the University of Pittsburgh.

Seung Han Ha received his Ph.D. in Biomedical Engineering (2009) from Korea University; he currently is a postdoctoral associate in Professor Kang Kim’s group at Pittsburgh.

Kang Kim is Assistant Professor of Medicine and Bioengineering at the University of Pittsburgh.

Anna A. Shvedova is Lead Research Physiologist at the Pathology and Physiology Research Branch, Health Effects Laboratory Division, National Institute for Occupational Safety and Health (NIOSH/CDC) and adjunct Professor at the Department of Physiology and Pharmacology, School of Medicine, West Virginia University.

Valerian E. Kagan is Professor and Vice Chair in the Department of Environmental and Occupational Health and Director of the Center for Free Radical and Antioxidant Health, at the University of Pittsburgh.

Alexander Star is Associate Professor of Chemistry also at Pittsburgh.

Three-dimensional rendering of X-ray crystal structures for (A) horseradish peroxidase (HRP) isoenzyme C and (B) myeloperoxidase (MPO) derived from human leukocytes. (C) Structure of the enzymes’ active site, a Fe(III) protoporphyrin IX heme group. (D) The catalytic peroxidase cycle for HRP and MPO. For MPO, Compound I is reduced directly to the resting state via conversion of Cl to hypochlorous acid in the neutrophils (red path). (A and B obtained from the Protein Data Bank, accession codes 1H5A and 3F9P, respectively. D adapted from ref. 19 © Elsevier, 2004).

TEM images of carboxylated single-walled carbon nanotubes (SWCNTs) incubated with horseradish peroxidase (HRP) and H2O2 at 4 °C (A) initially and (B) after 12 weeks. (C) Raman spectra depicting the change in graphitic structure of SWCNTs as a result of HRP/H2O2 catalyzed degradation at days 0 (blue), 7 (red), and 30 (cyan). (A and B adapted from 25 © American Chemical Society, 2008. C adapted from 27 © Royal Society of Chemistry, 2011).

(A) Schematic and transmission electron micrographs depicting the shortening of carboxylated multi-walled carbon nanotubes (MWCNTs) catalyzed by horseradish peroxidase (HRP) and H2O2. (B) High-resolution TEM images of carboxylated MWCNT before (left) and after (right) 60 days of enzymatic degradation, revealing a thinner nanotube after degradation. (C) Illustration of the mechanism of MWCNT degradation, in which carboxylated MWCNTs were oxidized and exfoliated layer-by-layer to yield pristine inner walls that were more resistant to HRP degradation. (A reproduced from 27 © Royal Society of Chemistry, 2011. B and C reproduced from 30 © American Chemical Society, 2011).

TEM micrographs of graphene oxide (GO) (A) initially and (B) after 10 days of incubation with horseradish peroxidase (HRP) and daily additions of 40 μM H2O2. (C) Raman spectra of GO and after 0 (black), 4 (red), and 20 (green) days of incubation with HRP/H2O2. (D) TEM image of reduced graphene oxide (RGO) after 20 days of identical incubation with HRP/H2O2. (A, B, C, and D reproduced from 31 © American Chemical Society, 2011).

(A) Photograph depicting dispersions of carboxylated SWCNTs alone (left) and incubated 24 h under different conditions: hMPO/H2O2/NaCl (center-left), hMPO/H2O2 (center-right), NaOCl (right). (B) Raman spectra (excitation, 633 nm) of carboxylated SWCNTs (black) and hMPO/H2O2-treated carboxylated SWCNTs at 10 (red), 15 (blue) and 24 (green) hours. (C) Vis-NIR spectra depicting the loss of M1 and S2 bands during the degradation of carboxylated SWCNTs in hMPO/H2O2 (OD, optical density). (D) Scanning electron micrographs illustrating biodegradation of carboxylated SWNTs (left) after 24 h under different conditions: hMPO/H2O2/NaCl (center-left), hMPO/H2O2 (center-right), NaOCl (right). (A, B, C, and D adapted from 10 © Macmillan Publishers Ltd., 2010).

Binding poses of (A) horseradish peroxidase (HRP) and (B) myeloperoxidase (MPO) on carboxylated SWCNTs calculated using molecular docking studies (AutoDock Vina). (C) Binding pose of HRP on graphene oxide. (A reproduced from 26 © American Chemical Society, 2009. B reproduced from 10 © Macmillan Publishers Ltd., 2010. C reproduced from 31 © American Chemical Society, 2011).

Raman spectra (excitation, 473 nm) recorded from different areas of neutrophils containing IgG–nanotubes at (A) 2 h and (B) 8 h. Inset depicts bright-field image of the neutrophils with engulfed IgG–nanotubes with cross-wires depicting location of laser spot. (C) D/G intensity ratios obtained from Raman spectroscopy on double-walled carbon nanotubes (DWCNTs) in PC3 cells at different time points. A trend of increasing defects on the walls of the DWCNTs was observed with the signal-to-noise ratio decreasing at 24 hours. (A–B reproduced from 10 © Macmillan Publishers Ltd., 2010. C reproduced from 39 © Wiley-VCH, 2010)

(A) Hyper-spectral images (CytoViva) of the lung tissue for both w/t and MPO k/o at day 1. The red areas are the pixels match the SWCNT spectral profiles. The scanned images were collected at 40× employing an Olympus BX-51 Microscope and a 100W Quartz-Halogen light source. Spectral data was captured with the CytoViva spectrophotometer utilizing an integrated CCD Camera. (B) Photoacoustic (PA) images and (C) statistical analysis (student’s t-test) of PA intensity averaged over the region in the box in (B) of a lung section from each group of animals at days 1 and 28. (D) Changes in the content of collagen (left) and the average thickness (right) of alveolar connective tissue in the lungs of w/t and MPO k/o mice at days 1 and 28 after pharyngeal aspiration of SWCNTs. (D reproduced from 11)

Chemical Methods to Oxidize Carbon Nanomaterials (CNMs)

The Longitudinal Unzipping of Single-walled Carbon Nanotubes (SWCNTs, reproduced from 28 © Macmillan Publishers Ltd., 2009)

The Effect of Functionalization on CNT Oxidation (Adapted from 34 © Elsevier, 2010)

Work Function (V) and Standard Potential (E0) of SWCNTs (Adapted from 35 © Macmillan Publishers Ltd., 2001 and 36 © American Chemical Society, 2002)

Redox Potentials of Peroxidases Used to Oxidize Carbon Nanomaterials

EnzymeStep on the Peroxidase CycleE0 (V)
HRP24Compound I/Compound II0.941
Compound II/Resting State0.960

MPO22Compound I/Resting State1.16
Compound I/Compound II1.35
Compound II/Resting State0.97

Functionalization of Carbon Nanomaterials and Their Reported ex vitro Degradation by Enzymatic Oxidation

NanomaterialFunctionalizationEnzymeOxidantReported DegradationReference
SWCNTPristine, unfunctionalizedHRPH2O2No[26]
CarboxylatedHRPH2O2Yes[2527]
CarboxylatedMPOH2O2, NaClYes[10,32]
Carboxylated/PhosphatidylserineMPOH2O2, NaClYes[10]

MWCNTPristine, unfunctionalizedHRPH2O2No[30]
CarboxylatedHRPH2O2Yes[27,30]
1,3-dipolar cycloadditionHRPH2O2Yes[33]
CONH(CH2CH2O)2CH2CH2NH3+HRPH2O2Yes[33]
Nitrogen-dopedHRPH2O2Yes[30]
CONH(CH2CH2O)2CH2CH2NH3+MPOH2O2, NaClYes[33]

GrapheneReduced Graphene Oxide (RGO)HRPH2O2No[31]
Graphene Oxide (GO)HRPH2O2Yes[31]