3
Carbon Nanomaterials for Optical Bioimaging and Phototherapy

Haifeng Dong and Yu Cao

Beijing Key Laboratory for Bioengineering and Sensing Technology, Research Center for Bioengineering and Sensing Technology, School of Chemistry & Biological Engineering, University of Science & Technology, Beijing, PR China

3.1 Introduction

Cancer is one of the most deadly diseases facing humanity [1]. The current standard cancer management includes stage determination, chemo/radiation therapy, and surgical resection. Despite great process has been made in the past few decades, early diagnosis and efficient treatment of cancer are still challenging to overcome [2]. Molecular imaging is a useful tool to monitor in vivo biochemical events and the development of nanomaterials as biomedical imaging agents is a very promising method to obtain detailed images in living systems [3, 4]. It is beneficial for the researchers to follow the distribution of the drug inside the organism and gives further hints for the optimization of disease treatment to combine the drug delivery features with imaging techniques [5, 6]. The combination of diagnostic tools such as optical imaging with therapeutical approaches such as chemotherapy and phototherapy gives rise to promising theranostic nanomaterials [79]. As the forefront of theranostic nanomaterials for cancer therapy, the unique optical properties of carbon nanomaterials such as fullerenes [10], nanodiamonds (NDs) [11, 12], carbon nanotubes (CNTs) [13, 14], graphene and its derivatives [15, 16], and carbon quantum dots (CQDs) [17] (Figure 3.1) have inspired extensive studies due to their great potential applications in the field of optical bioimaging analysis and phototherapies [18, 19].

Image described by caption.

Figure 3.1 Carbon nanomaterials including fullerene, carbon nanotube, graphene, carbon dot, and nanodiamond.

Source: Reprinted with permission from Ref. [18].

In this chapter we will cover the recent progress in optical biological imaging analysis and phototherapies using carbon nanomaterials.

3.2 Surface Functionalization of Carbon Nanomaterials

Since most pristine carbon nanomaterials are highly hydrophobic due to the sp2 carbon nanostructures, proper functionalization is necessary to improve their water solubility and biocompatibility before their biomedical applications. Generally, there are two primary strategies of functionalization: covalent or noncovalent. Covalent modifications usually involve introduction of hydrophilic functional groups (e.g. hydroxyl groups, carboxyl groups, and amino groups) for the further conjugation of protecting polymers (such as polyethylene glycol, PEG), targeting ligand, or drug/gene cargos [2024]. The covalently functionalized carbonaceous nanomaterials are usually stable. However, the limitation lies in that this type of functionalization method inevitably destroys partial material structure, causing the loss of certain intrinsic properties (e.g. photothermal capacities). Compared with covalent functionalization, the reaction condition of noncovalent functionalization is comparatively mild, which involves coating the carbonaceous nanomaterials with amphiphilic molecules [25]. The hydrophobic motifs of the amphiphilic molecules could be anchored onto the material's surface with the hydrophilic ends extending to the aqueous solution and maintaining the stability of the whole material. Noncovalent interactions include electrostatic forces, π–π interactions, hydrogen bonding, and van der Waals forces. However, lower stability of noncovalent conjugates is the major concern for this type of functionalization. A lot of thought about stability and design of the nanomaterial must be taken into account before choosing an appropriate functionalization method for any carbonaceous nanomaterial. A balance between stability and structural integrity must be maintained before any further biomedical applications.

Different carbon nanomaterials require different strategies of surface functionalization to make them soluble in aqueous environment and compatible with cells and tissues. Fullerenes are typically covalently functionalized through chemical reactions directly with the carbon atoms in the sp2 carbon shell [26], and a library of standard chemical reactions have been developed for fullerene chemistry. CQDs and graphene quantum dots (GQDs) are by nature rich in ─OH and ─COOH functional groups, which can easily form hydrogen bonds with water molecules and thus endow them with good solubility in aqueous environment; nonetheless, it is still desirable to further functionalize them with PEG or other functional groups to increase biocompatibility. CNTs [21, 27] and graphene [28, 29], both of which feature continuous graphitic honeycomb structures expanding over submicrometer to micrometer scales, can be either covalently or noncovalently functionalized to impart water solubility, depending on the need for specific biological applications. Nanodiamond [22], on the other hand, is similar to CNTs and graphene that both noncovalent and covalent functionalizations have been reported to increase their water solubility and biocompatibility.

3.3 Carbon Nanomaterials for Optical Imaging

Many organic fluorophores absorb and emit light in the visible spectral range, which can have some drawbacks. In complex biological systems like cells, or especially in living organisms, the absorption and autofluorescence of the tissue and body fluids are a major problem for optical imaging. The absorption reduces the transmission of the excitation light, and also the emitted fluorescence signal is significantly weakened, or even completely quenched. To overcome this issue, researchers have found that in the near‐infrared (NIR) region between 650 and 950 nm, the so‐called NIR I, hemoglobin and water, as the two main absorbers, have low molar extinction coefficients [30]. NIR light can penetrate the tissue, and the use of NIR‐emitting dyes with emission maxima in the NIR I as fluorescent labels allows for deep tissue imaging. Another optical window, the NIR II, has been identified later in the spectral range between 1000 and 1350 nm [31].

Several spectroscopic techniques can be utilized for biological in vitro and in vivo imaging. One of the most common techniques is fluorescence imaging. Carbon nanomaterials can be used for their intrinsic fluorescence properties or can be tagged with fluorescent molecules.

3.3.1 Intrinsic Fluorescence of Carbon Nanomaterials

A fullerene nanoparticle is made of a closed shell of graphene that contains conjugated double bonds [32]. Therefore, electronic transitions are expected to take place between the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) of the large π‐conjugated system, corresponding to strong absorption mostly in the ultraviolet (UV) region (<400 nm), which should be followed by fluorescence emission in the visible window. However, the fluorescence quantum yield of unmodified, pristine fullerene is very low, on the order of 10−5 –10−4 , and the photoluminescence decreases at increasing temperature with very weak emission at room temperature due to nonradiative recombination of excited nanoparticles coupled with phonons [33].

Semiconducting single‐walled carbon nanotubes (SWCNTs) show a structure dependent fluorescence in the biological important NIR II window that have been widely studied [34, 35]. The band gap fluorescence of semiconducting SWCNTs originates from light absorption at a photon energy of E22, creating an electron–hole pair. This initial excitation is followed by a nonradiative relaxation of the electron to the conduction band and then the fluorescence event with a photon energy of E11. The fluorescence quantum yields of single‐walled nanotubes (SWNTs) are low, typically lower than 0.01 for macroscopic samples of SWCNT [36, 37], and show a strong dependency on environmental factors. Recent advance in chemical doping of SWCNTs through oxygen or sp3 incorporation has led to largely increased fluorescence quantum yield [38]. Also in the visible region, fluorescence of well‐dispersed CNTs was reported, which showed a large dependence on the type of the CNT functionalization [39].

GQDs and CQDs are another example of fluorescent carbon nanomaterials, which are widely researched for biological imaging applications. The mechanisms of the GQD fluorescence are currently under investigation and still, the physical origin of the fluorescence seems not to be fully understood [40, 41]. However, structurally perfect graphene sheet is a zero‐band gap semiconductor with the conduction band and valence band touching at the K points and thus show no luminescence upon photo‐excitation. One common strategy to create fluorescent graphene derivatives is the introduction of defect sites to open the band gap predictably due to quantum confinement and edge effects. Besides, the modification of heteroatoms (such as O, N, P, and S) in the GQDs and CQDs also plays a crucial role in determining the nonradiative decay rate and even the band gap. An interesting feature of GQDs is their excitation wavelength dependent emission. In line with modulation of the excitation wavelength, the emission wavelength undergoes a shift.

The distinctive optical properties of graphene and its derivatives allow them for biomedical imaging in vitro and in vivo [42]. Although pristine graphene is a zero‐gap semiconductor as aforementioned, due to the highly heterogeneous chemical and electronic structures of graphene oxide (GO), strong fluorescence emission has been observed from GO in a broad spectrum, ranging from the UV to NIR, the exact mechanism of which remains to be fully elucidated and has largely been believed to arise from the electronic transition between the nonoxidized, pristine sp2 carbon domain and the oxidized boundaries of GO sheet.

Due to the lack of sp2 graphitic carbon, the intrinsic fluorescence emission of nanodiamond can be led back to the presence of nitrogen‐vacancy (N‐V) centers in the structure of the nanodiamond. Its fluorescence is very robust against photo bleaching, making nanodiamond an ideal fluorescence label for long‐term tracking of biomarkers and cells [43, 44].

3.3.2 Imaging Utilizing Intrinsic Fluorescence Features of Carbon Nanomaterials

Fullerene and its derivatives have relatively less bright fluorescence compared to other carbon nanomaterials, and the photodynamic property of fullerene also raises concern on generating active oxygen species after photoexcitation and affecting the cell viability. Nonetheless, there are a handful of reports on using C60 and C70 as fluorescent biomarkers. Prasad et al. have studied the internalization of 750 nm fluorescence emitting C60 aggregates into the cytoplasm of MCF10A cells, which are normal human mammary epithelial cells, with fluorescence costaining of F‐actin using fluoresceinisothiocyanate (FITC) phalloidin [45]. Chung et al. performed the fluorescence imaging of HeLa cells in different excitation/emission channels with distinct colors for C60‐tetraethylene glycol (C60‐TEG) proceeded the tunable fluorescence color depend on the concentration of C60‐TEG in the solution [46]. Lee et al. have reported on using hyaluronated C60 particles for tumor targeting and in vivo fluorescence imaging on tumor‐bearing mice [47]. The hyaluronate‐C60 bioconjugate allows targeted in vivo tumor fluorescence imaging by detecting the deep red fluorescence of hyaluronated C60 at 710 nm under an excitation of 635 nm (Figure 3.2).

Image described by caption.

Figure 3.2 Noninvasive photoluminescent (635/710) imaging of nude mice harboring HCT‐116 or KB tumors. HA‐F1 was intravenously injected into tumor‐bearing nude mice. Photoluminescent images were obtained for four hours after injection. Tumors are indicated by the white arrows. (See color plate section for the color representation of this figure.)

Source: Reprinted with permission from Ref. [47].

GQDs and CQDs are widely applied for optical bioimaging [48, 49]. As early as 2011, Zhu et al. synthesized GQDs through the hydrothermal method from GO, which had a fluorescence quantum yield as high as 0.11, showing excitation‐dependent and solvent‐dependent fluorescence in the visible region (Figure 3.3a) [50]. The results demonstrated that GQDs could be used for fluorescence imaging of cancer cells. Since 2014, high‐quality studies on the preparation of CQDs and GQDs and their application for bioimaging have been published [5254]. Similar to GQDs, the emission wavelengths of the carbon dots (CDs) shifted, depending on the excitation wavelengths. GQDs and CQDs are also applicable for two‐photon bioimaging [55]. In a two‐photon excitation process, the fluorophore absorbs two photons simultaneously and is excited to a higher energy state, after which the absorbed energy is released by emitting only one photon with a shorter wavelength than the absorbed photons as the molecule returns to the ground state. The group of Sun first observed one‐ and two‐photon fluorescence in CQDs upon excitation with a 458 nm laser or a femtosecond pulsed laser at 800 nm, respectively [51]. When the CQDs were internalized by MCF‐7 cells irradiating with the pulsed 800 nm laser, green two‐photon luminescence was observed (Figure 3.3b). Also, GQDs were successfully used for two‐photon‐induced fluorescence for cellular and deep tissue imaging [56].

Image described by caption.

Figure 3.3 (a) Cellular imaging of GQDs under 405 and 488 nm excitation, respectively; (b) Representative two‐photon luminescence image (800 nm excitation) of human breast cancer MCF‐7 cells with internalized CDs. (See color plate section for the color representation of this figure.)

Source: (a) Reprinted with permission from Ref. [50]. (b) Reprinted with permission from Ref. [51].

Fluorescent CNTs have been widely studied as in vitro and in vivo imaging agents. The first report of NIR‐emitted SWCNTs for cellular imaging dates back to 2004, when Weisman et al. incubated macrophage cells with SWCNTs [34]. The SWCNT NIR emission remained intact upon cellular uptake but underwent a slight bathochromic shift. This study rendered the microscopic detection of fluorescent SWCNTs inside of cells as a powerful method to investigate the interactions of CNTs with cells. Neves et al. investigated the internalization and fate of CNTs inside cells in detail [57]. The results revealed that an endocytic pathway was involved in the internalization of RNA‐wrapped, oxidized double‐walled CNTs (oxDWNT‐RNA) (Figure 3.4a,b). The nanotubes were found in clathrin‐coated vesicles, after which they appear to be sorted in early endosomes, followed by vesicular maturation, become located in lysosomes. The first in vivo research was proceeded on the larvae of Drosophila melanogaster which were fed with food‐containing SWCNTs [58]. The NIR fluorescence signal could be detected in the living larvae and was used for imaging (Figure 3.4c–e). Dai et al. for the first time studied the NIR fluorescence of SWCNTs on mice in 2009 [59]. They utilized different kinds of NIR fluorescent phospholipid‐polyethylene glycol (PL‐PEG) coated SWCNTs for targeted cell imaging by attaching an arginine‐glycine‐aspartic acid peptide, also known as an Arg‐Gly‐Asp, or RGD peptide. In vivo studies were carried out with nude mice and tumor bearing LS174T mice. After injection, the PL‐PEG coated SWCNTs showed bright NIR fluorescence imaging with good contrast (Figure 3.4f,g). The SWCNTs could be identified in the vasculature under the skin and also in deeper organs such as liver and spleen. With time, the SWCNTs were cleared from the body. As discussed above, SWCNTs have great promise for use as effective biomarkers for in vitro and in vivo imaging due to their excellent properties such as low cytotoxicity, high photostability, absence of quenching, and photo bleaching if functionalized properly.

Image described by caption.

Figure 3.4 (a, b) Subcellular distribution of CNT complexes in PC3 cells. Cells were incubated with 30 g ml−1 of oxDWNT‐RNA‐FS for two hours; (c–e) SWNT NIR emission showing accumulation in the dorsal vessel. Green fluorescence from green fluorescent protein (GFP) expressed exclusively in the dorsal vessel is shown in panel (c), and NIR fluorescence from nanotubes is shown in panel (d) (false colored in red). (e) Overlay of these two images on the corresponding bright field image, demonstrating that the SWNTs lie within the lumen of the vessel. Scale bars are 25 μm. NIR photoluminescence images of nude mice treated with 200 μl of 17 mg ml−1 exchange‐SWNTs (f) and 200 μl of 260 mg ml−1 direct‐SWNTs (g). (See color plate section for the color representation of this figure.)

Source: (a, b) Reprinted with permission from Ref. [57]. (c–e) Reprinted with permission from Ref. [58]. (f, g) Reprinted with permission from Ref. [59].

Graphene and its derivatives have found applications in bioimaging owing to their intrinsic optical properties. The Dai’s group for the first time utilized the NIR photoluminescence of nGO‐PEG for cell imaging [60]. It was found that GO and nGO‐PEG exhibited fluorescence broadly from visible to NIR‐I window could be used for cell imaging with little background. Although the quantum yield of nGO‐PEG fluorescence was rather low, cell imaging using its intrinsic fluorescence has been successfully realized owing to the extremely low auto‐fluorescence background of biological tissues in the NIR‐I window.

Nanodiamonds have been used for bioimaging for some time and the emission wavelength of nanodiamonds is located in the visible region, with characteristic green to red emission between 550 and 800 nm [61]. Chang et al. reported the application of fluorescent nanodiamonds as nonbleaching imaging agents for human 293T kidney cells with no observed cytotoxicity in early 2005 [62]. After that, so many researches have been studied signaling the huge potential of nanodiamonds as high‐quality imaging agents. In consecutive work, the uptake mechanism of nanodiamonds was further studied showing that nanodiamonds entered the cells mainly by endocytosis, but also on a clathrin‐mediated pathway [63]. While larger nanodiamond nanoparticles localized in vesicles, small nanodiamonds appeared to be in the cytoplasm. Nanodiamonds were also used as fluorescent probes for in vivo studies. The interaction of two different kinds of nanodiamonds (with 5 and 100 nm diameter) with the ciliated eukaryotic unicellular organisms Paramecium caudatum and Tetrahymena thermophila was studied. The nanodiamonds revealed low toxicity and excellent fluorescence properties. The effects of fluorescent nanodiamonds were also investigated on Caenorhabditis elegans and zebrafish [64, 65]. It was found that nanodiamonds could be delivered to the embryos of the next generation providing great capacity for fluorescent in vivo imaging (Figure 3.5). Cheng et al. investigated the long‐term stability and biocompatibility of 100 nm diameter nanodiamonds on mice [66]. The in vivo whole‐body imaging and ex vivo examination of various body tissues suggested that no toxic effects of the nanodiamonds on the mice, rendering promising application of nanodiamonds for in vivo imaging.

Image described by caption.

Figure 3.5 Differential interference contrast (DIC) and stacked confocal fluorescence images of bovine serum albumin (BSA)‐conjugated fluorescent nanodiamonds (FNDs) in zebrafish embryos at 4.7 hpf (a) and ~60 hpf (b). The corresponding control experiments were performed without FND injection to the zebrafish embryos. The developmental ages are 4.3 hpf (c) and ~60 hpf (d), respectively. (See color plate section for the color representation of this figure.)

Source: Reprinted with permission from Ref. [65].

3.3.3 Imaging with Fluorescently Labeled Carbon Nanomaterials

Due to their excellent chemical and physical properties like a high surface area, CNTs are widely employed in biomedical applications such as drug and gene delivery. Fluorescein derivatives are widely used as fluorescent labels in combination with CNTs. In 2004, Prato et al. used fluorescein‐tagged SWCNTs to study the internalization of CNTs by human 3T6 and murine 3T3 fibroblasts [67]. Then the cellular uptake mechanisms of functionalized CNTs were studied in detail. It was shown that covalently functionalized CNTs, bearing different fluorescein‐based labels for imaging in addition to other functional groups, were taken up by a variety of different cell lines. The group of Dai exhibited one example for noncovalently functionalized fluorescein labeled SWCNTs [68]. Fluorescein functionalized PEG was physiosorbed onto the surface of SWCNTs, affording stable aqueous suspensions. The fluorescein labeled DNA was also used to functionalize SWCNTs noncovalently to monitor the cellular uptake of the carbon nanomaterial [69].

The low quantum yield of GO fluorescence and the auto‐fluorescence interference of biological tissues have limited the applications of GO in bioimaging especially in vivo animal imaging. Therefore, a number of groups used external fluorescent dyes to label GO‐PEG for in vitro and in vivo imaging. Liu et al. labeled nGO‐PEG with a NIR dye, Cy7, by conjugating the dye molecule at the PEG terminal via covalent modification to avoid significant fluorescent quenching [70]. Increasing fluorescence signals were found in the tumor over time, indicating efficient tumor passive targeting of nGO‐PEG in several different xenograft tumor models (Figure 3.6). In order to avoid fluorescence quenching or photo bleaching, radiolabeling is a much more sensitive method for in vivo bioimaging. Liu’s group developed a method to label nGO‐with 125 I by anchoring iodine atoms on the defects and edges of GO [71]. Then they fabricated 64 Cu labeled nGO‐PEG for in vivo positron emission tomography (PET) imaging [72]. Furthermore, by noncovalent modifying with DNA or miRNA detection probe, graphene and its derivatives can be used for in vitro or intracellular miRNA imaging.

Image described by caption.

Figure 3.6 Spectrally unmixed in vivo fluorescence images of 4T1 tumor‐bearing Balb/c mice, KB, and U87MG tumor‐bearing nude mice at different time points post injection of NGS‐PEG‐Cy7. Mouse autofluorescence was removed by spectral unmixing in the above images. High tumor uptake of NGSPEG‐Cy7 was observed for all of the three tumor models. Hairs on Balb/c mice were removed before fluorescence imaging.

Source: Reprinted with permission from Ref. [70].

Besides the intrinsic fluorescence of nanodiamonds, some reports focus on nanodiamonds functionalized with fluorescein and other dyes for bioimaging. In 2008, the group of Hwang modified fluorescein on the surface of the nanodiamonds for cell imaging [73]. The bright green luminescence of the fluorescein dye was observed inside the cell. Zhang et al. developed a nanodiamond‐based drug delivery system by functionalizing with an anti−epidermal growth factor receptor (EGRF) antibody for targeting and a fluorescein labeled drug‐oligonucleotide conjugate for therapy [74]. The fluorescence of the fluorescein could trace the nanodiamonds inside the cells. As a kind of commonly used anticancer drugs, doxorubicin (DOX) with fluorescence was loaded on nanodiamonds to investigated in vitro and in vivo cancer therapy efficiency [75].

3.4 Carbon Nanomaterials for Phototherapies of Cancer

Chemotherapy and radiotherapy are major therapeutic approaches for the treatment of a wide variety of cancers in recent year. However, one of major disadvantages of chemotherapy and radiotherapy is their limited specificity to cancer cells and could lead to undesired side effects to normal tissues and organs. Phototherapies, mainly including photothermal therapy (PTT) and photodynamic therapy (PDT), are able to destruct cancer cells upon specific light irradiation [6, 7678]. With the help of nanotechnology, phototherapeutic nanoagents could specifically target cancer via either passive or active tumor targeting. Therefore, phototherapies exhibit remarkable advantages in terms of enhancing cancer killing specificity and reducing side effects, in comparison to conventional cancer therapies. In the past few years, phototherapies based on the unique optical and chemical properties of carbon nanomaterials have aroused increasing interest [2, 16, 79, 80].

3.4.1 Photothermal Therapy

PTT uses external light‐induced hyperthermia on malignant tissues to kill abnormal cells while avoiding damage to healthy tissue [81, 82]. Carbon nanomaterials such as CNTs and graphene selectively accumulate in tumors through the enhanced permeability and retention (EPR) effect as well as demonstrate extremely high levels of intrinsic absorbance in the biological transparency windows located within the NIR window (750–1700 nm) [16, 83]. Additionally, the strong fluorescence of carbon nanomaterials with emission wavelengths ranging from visible to NIR‐II window allows for simultaneous imaging and treatment. Compared to other nanomaterials including various gold nanostructures [8486], copper sulphide (CuS) nanoparticles [8789], and organic nanoparticles [90], the proper surface modified carbon nanomaterials can circulate in vivo with longer half‐lives.

SWCNTs were the first carbon nanomaterial used for PTT. In 2009, Choi et al. reported photothermal ablation of human epidermoid mouth carcinoma KB tumor in mouse intratumorally injected with PEGylated SWCNTs using 808 nm laser irradiation [91]. A significantly reduced tumor volume was found for the treated mice compared to the control. In 2010, Dai’s group reported the use of SWCNTs noncovalently functionalized with PL‐PEG and C18‐PMH‐PEG to achieve a high passive tumor uptake based on the EPR effect for efficient PTT [92]. This was the first work to image tumors in the NIR‐II window with SWCNTs and the first systemic injection of these nanomaterials for PTT. Since then, a lot of studies have been done for PTT to decrease the injection dose and irradiating power density [93, 94].

Besides SWCNTs, graphene, and its derivatives have also been successfully applied for PTT owing to their strong absorbance in the NIR window [95]. In 2010, Liu et al. for the first time studied the behaviors of PEGylated nano‐GO in mice after intravenous (i.v.) injection by a fluorescent labeling and imaging method and revealed the EPR effect of cancerous tumors [70]. Owing to the high uptake inside the tumor, the photothermal agent nGO‐PEG achieved efficient tumor ablating with 100% of tumors eliminated under 808 nm laser irradiation at a power density of 2 W cm−2 for five minutes (Figure 3.7a). While the injected doses (20 mg kg−1 ) and NIR laser densities (2 W cm−2 ) are comparable to that of other photothermal agents such as gold nanomaterials, showing that nGO‐PEG appeared to be an excellent in vivo tumor NIR PTT agent without exhibiting noticeable toxicity. It has been well known that the reduction of GO could lead to dramatically enhanced optical absorbance in the NIR region. In order to reduce the requisite dosage and laser power density for PTT, Dai et al. reported that ultra‐small reduced nanographene oxide (nRGO) could be prepared by the reduction of PEGylated GO, indicating a significant increase of absorbance in the visible to NIR window (Figure 3.7b) [96]. By further coating with PEGylated phospholipid and conjugating with an RGD‐targeted peptide, the nRGO‐PEG could act as a highly efficient targeted photothermal agent for in vitro selective cancer cell ablation.

Image described by caption.

Figure 3.7 (a) Representative photos of tumors on mice after various treatments indicated. The laser‐irradiated tumor on the nanographene sheet (NGS)‐injected mouse was completely destructed; (b) Thermal images of vials containing pellets of control nontreated U87MG cells, cells treated by nano‐rGO‐RGD, and cells treated by nano‐rGO‐RAD, respectively, after eight minutes of irradiation with an 808 nm laser at a power of 15.3 W cm−2 . (See color plate section for the color representation of this figure.)

Source: (a) Reprinted with permission from Ref. [70]. (b) Reprinted with permission from Ref. [96].

3.4.2 Photodynamic Therapy

Different from PTT, PDT relies on reactive oxygen species (ROS) such as single oxygen (1 O2) produced by photosensitizer (PS) molecules to kill cancer cells under the irradiation of light with appropriate wavelengths [9799]. Three pivotal factors account for PDT therapeutic efficiency: (i) PS photo‐transferring efficiency, (ii) light of the wavelength that activates the PS, and (iii) molecular oxygen. While less research has been conducted on using carbon nanomaterials as PS for PDT, the synergistic effects of combining carbon nanomaterials with PS significantly boost the efficacy of cancer treatment. PS can be noncovalently loaded onto the aromatic surface of carbon nanomaterials coated in biocompatible polymers through π–π stacking for activation after laser excitation.

Fullerenes such as C60 and C70 have been found as PSs with a high efficiency of ROS generation, allowing for efficient PDT using fullerenes without loading additional photosensitizing molecules [100]. Hamblin et al. reported PDT using N‐methylpyrrolidinium‐fullerene for the treatment of colon adenocarcinoma by intraperitoneal injection of the functionalized fullerene into the abdomens of the tumor‐bearing mice and found significantly slowed tumor growth compared to the control group [101]. The unique property of fullerenes in the efficient conversion from photons to ROS has made fullerene derivatives competitive PSs for PDT applications in in vivo animal and even preclinical settings [102].

CNTs can either act as PSs themselves or load exogenous PSs for PDT. Imahori et al. found semiconducting SWCNTs could generate ROS more efficiently than metallic SWCNTs, opening up the possibility of using chemically separated, pure semiconducting SWCNTs as PSs for PDT [103]. Zhang et al. have demonstrated that photodynamic effects of SWCNTs highly depend on the modification method of the nanotubes, and effective in vivo tumor destruction has been realized based on PDT using polyethyleneimine (PEI) and polyvinylpyrrolidone (PVP) modified SWCNTs [104]. SWCNTs can also be used as nanocarriers to deliver PSs to target tissue such as the tumor, owing to the high accumulation of SWCNTs inside tumor tissue based on the EPR effect. Lee group reported PEGylated SWCNTs loaded with Chlorin e6 (Ce6) could be applied for PDT‐based tumor treatment in mice [105].

By taking advantage of the ultrahigh loading of molecules through π–π stacking, nano‐GO has been used to deliver PSs in vitro at high concentrations [106]. The first study of the graphene‐based PDT was reported by Shi et al. In that work, zinc phthalocyanine (ZnPc), a widely used PS molecule, was loaded on the surface of nGO‐PEG via p–p stacking and hydrophobic interactions. The obtained nGO‐PEG‐ZnPC exhibited significant cytotoxicity toward cancer cells under Xe light irradiation. In a later work, Huang et al. reported that folic acid‐conjugated GO could load Ce6 for targeted delivery to fluorescence resonance (FR) positive cancer cells and achieved effective cancer cell photodynamic destruction using a 633 nm He─Ne laser irradiation [107].

Recently, GQDs have been reported as PS to efficiently generate singlet oxygen (1 O2) for PDT [108, 109]. Ge et al. presented GQDs as a new PDT agent to produce 1 O2 via a multistate sensitization process, resulting in a quantum yield of about 1.3 [110]. The as‐prepared GQDs exhibit a combination of properties, including broad absorption from the visible to the NIR, deep‐red emission, good aqueous dispersibility, high photo‐ and pH stability, and favorable biocompatibility, which enable them to act as a multifunctional nanoplatform for the simultaneous imaging and highly efficient in vivo PDT of cancer (Figure 3.8).

Image described by caption.

Figure 3.8 Dose‐dependent PDT effects of the cell viability of HeLa cells: (a) GQDs in the concentration range 0.036–1.8 μM; (b) PpIX in the concentration range 0.36–18 μM; (c) Photographs of mice after various treatments on the 1st, 9th, 17th, and 25th day. (PDT: GQDs þ light irradiation; C1: GQDs only; C2: light irradiation only.)

Source: Reprinted with permission from Ref. [110].

3.5 Conclusions and Outlook

Carbon nanomaterials with exciting chemical, mechanical, and optical properties have had a tremendous impact on the development of theranostic methods for in vitro and in vivo biomedical applications, including imaging, drug delivery, and phototherapies. Although all made of the same chemical element, these nanomaterials with different allotropic forms of carbon exhibit distinct properties and behaviors, depending on how the carbon atoms are bonded to form the larger structures on the nanoscale and on the size of the nanostructures. With a library of well‐established surface functionalization and passivation methodologies, it is now possible to make water‐soluble and biocompatible carbon nanomaterials with minimum in vitro and in vivo toxicity, and to use these nanostructures for a myriad of biomedical imaging and therapeutic applications.

To achieve effective in vivo live animal imaging, PTT and PDT, the light penetration depth is the main challenge to overcome. Carbon nanomaterials like SWCNTs and GOs processed NIR fluorescence have shown great potential for deep tissue imaging and phototherapy. But it still remains highly challenging for deeper imaging and therapy to centimeters in a living organism. Several possible directions for deeper fluorescence imaging and phototherapy are given: first, carbon nanomaterials can be engineered to have even longer fluorescence emission wavelength to afford deeper in vivo fluorescence imaging and phototherapy; second, improving the fluorescence quantum yields for deeper in vivo imaging and increasing the photothermal conversion efficiency for more efficient PTT; third, combining other diagnostic tools and therapeutic approaches for synergetic diagnostic and therapy.

The focus of previous research efforts was on CNTs, nanodiamonds and graphene‐based nanomaterials. However, the future clearly lies in the development of sophisticated, multifunctional theranostic nanomaterials with properties for specific purposes. Other carbon nanomaterials like carbon nanohorns (CNHs), carbon nanoonions (CNOs) and graphdiyne, are highly promising for future developments. We believed that with innumerable exciting opportunities existing down the road, more environmentally friendly and human friendly products and techniques of carbon nanomaterials would be developed in the future.

References

  1. 1 Siegel, R.L., Miller, K.D., and Jemal, A. (2016). Cancer statistics, 2016. CA Cancer J. Clin. 66 (1): 7–30.
  2. 2 Chen, D., Dougherty, C.A., Zhu, K. et al. (2015). Theranostic applications of carbon nanomaterials in cancer: focus on imaging and cargo delivery. J. Controlled Release 210: 230–245.
  3. 3 Koo, H., Huh, M.S., Ryu, J.H. et al. (2011). Nanoprobes for biomedical imaging in living systems. Nano Today 6 (2): 204–220.
  4. 4 Wen, J., Xu, Y., Li, H. et al. (2015). Recent applications of carbon nanomaterials in fluorescence biosensing and bioimaging. Chem. Commun. (Cambridge) 51 (57): 11346–11358.
  5. 5 Etrych, T., Lucas, H., Janouskova, O. et al. (2016). Fluorescence optical imaging in anticancer drug delivery. J. Controlled Release 226: 168–181.
  6. 6 Parveen, S., Misra, R., and Sahoo, S.K. (2012). Nanoparticles: a boon to drug delivery, therapeutics, diagnostics and imaging. Nanomedicine 8 (2): 147–166.
  7. 7 Ma, X., Zhao, Y., and Liang, X.J. (2011). Theranostic nanoparticles engineered for clinic and pharmaceutics. Acc. Chem. Res. 44 (10): 1114–1122.
  8. 8 Ranjan, S., Jayakumar, M.K., and Zhang, Y. (2015). Luminescent lanthanide nanomaterials: an emerging tool for theranostic applications. Nanomedicine (London) 10 (9): 1477–1491.
  9. 9 Jeong, E.H., Jung, G., Hong, C.A. et al. (2014). Gold nanoparticle (AuNP)‐based drug delivery and molecular imaging for biomedical applications. Arch. Pharmacal Res. 37 (1): 53–59.
  10. 10 Bosi, S., Da Ros, T., Spalluto, G. et al. (2003). Fullerene derivatives: an attractive tool for biological applications. Eur. J. Med. Chem. 38 (11–12): 913–923.
  11. 11 Waddington, D.E.J., Sarracanie, M., Zhang, H. et al. (2017). Nanodiamond‐enhanced MRI via in situ hyperpolarization. Nat. Commun. 8: 15118.
  12. 12 Vaijayanthimala, V., Lee, D.K., Kim, S.V. et al. (2015). Nanodiamond‐mediated drug delivery and imaging: challenges and opportunities. Expert Opin. Drug Deliv. 12 (5): 735–749.
  13. 13 Bianco, A., Kostarelos, K., and Prato, M. (2005). Applications of carbon nanotubes in drug delivery. Curr. Opin. Chem. Biol. 9 (6): 674–679.
  14. 14 Ji, S.R., Liu, C., Zhang, B. et al. (2010). Carbon nanotubes in cancer diagnosis and therapy. Biochim. Biophys. Acta 1806 (1): 29–35.
  15. 15 Yang, G., Zhu, C., Du, D. et al. (2015). Graphene‐like two‐dimensional layered nanomaterials: applications in biosensors and nanomedicine. Nanoscale 7 (34): 14217–14231.
  16. 16 Yang, K., Feng, L., Shi, X. et al. (2013). Nano‐graphene in biomedicine: theranostic applications. Chem. Soc. Rev. 42 (2): 530–547.
  17. 17 Lim, S.Y., Shen, W., and Gao, Z. (2015). Carbon quantum dots and their applications. Chem. Soc. Rev. 44 (1): 362–381.
  18. 18 Hong, G., Diao, S., Antaris, A.L. et al. (2015). Carbon nanomaterials for biological imaging and nanomedicinal therapy. Chem. Rev. 115 (19): 10816–10906.
  19. 19 Yang, W., Ratinac, K.R., Ringer, S.P. et al. (2010). Carbon nanomaterials in biosensors: should you use nanotubes or graphene? Angew. Chem. Int. Ed. 49 (12): 2114–2138.
  20. 20 Bai, H., Li, C., and Shi, G. (2011). Functional composite materials based on chemically converted graphene. Adv. Mater. 23 (9): 1089–1115.
  21. 21 Hirsch, A. (2002). Functionalization of single‐walled carbon nanotubes. Angew. Chem. Int. Ed. 41 (11): 1853.
  22. 22 Lai, L. and Barnard, A.S. (2015). Functionalized nanodiamonds for biological and medical applications. J. Nanosci. Nanotechnol. 15 (2): 989–999.
  23. 23 Shim, M., Shi Kam, N.W., Chen, R.J. et al. (2002). Functionalization of carbon nanotubes for biocompatibility and biomolecular recognition. Nano Lett. 2 (4): 285–288.
  24. 24 Prencipe, G., Tabakman, S.M., Welsher, K. et al. (2009). PEG branched polymer for functionalization of nanomaterials with ultralong blood circulation. J. Am. Chem. Soc. 131 (13): 4783–4787.
  25. 25 Chen, R.J., Bangsaruntip, S., Drouvalakis, K.A. et al. (2003). Noncovalent functionalization of carbon nanotubes for highly specific electronic biosensors. Proc. Natl. Acad. Sci. U. S. A. 100 (9): 4984–4989.
  26. 26 Langa, F., de la Cruz, P., Espíldora, E. et al. (2000). Fullerene chemistry under microwave irradiation. Carbon 38 (11–12): 1641–1646.
  27. 27 Tasis, D., Tagmatarchis, N., Bianco, A. et al. (2006). Chemistry of carbon nanotubes. Chem. Rev. 106 (3): 1105–1136.
  28. 28 Zhu, S., Zhang, J., Tang, S. et al. (2012). Surface chemistry routes to modulate the photoluminescence of graphene quantum dots: from fluorescence mechanism to up‐conversion bioimaging applications. Adv. Funct. Mater. 22 (22): 4732–4740.
  29. 29 Wang, Y., Li, Z., Wang, J. et al. (2011). Graphene and graphene oxide: biofunctionalization and applications in biotechnology. Trends Biotechnol. 29 (5): 205–212.
  30. 30 Weissleder, R. (2001). A clearer vision for in vivo imaging. Nat. Biotechnol. 19 (4): 316–317.
  31. 31 Smith, A.M., Mancini, M.C., and Nie, S. (2009). Bioimaging: second window for in vivo imaging. Nat. Nanotechnol. 4 (11): 710–711.
  32. 32 Reuther, U. and Hirsch, A. (2000). Synthesis, properties and chemistry of Aza[60]fullerene. Carbon 38 (11–12): 1539–1549.
  33. 33 Capozzi, V., Casamassima, G., Lorusso, G.F. et al. (1996). Optical spectra and photoluminescence of C60 thin films. Solid State Commun. 98 (9): 853–858.
  34. 34 Cherukuri, P., Bachilo, S.M., Litovsky, S.H. et al. (2004). Near‐infrared fluorescence microscopy of single‐walled carbon nanotubes in phagocytic cells. J. Am. Chem. Soc. 126 (48): 15638–15639.
  35. 35 Heller, D.A., Baik, S.T., Eurell, E. et al. (2005). Single‐walled carbon nanotube spectroscopy in live cells: towards long‐term labels and optical sensors. Adv. Mater. 17 (23): 2793–2799.
  36. 36 Carlson, L.J., Maccagnano, S.E., Zheng, M. et al. (2007). Fluorescence efficiency of individual carbon nanotubes. Nano Lett. 7 (12): 3698–3703.
  37. 37 Crochet, J., Clemens, M., and Hertel, T. (2007). Quantum yield heterogeneities of aqueous single‐wall carbon nanotube suspensions. J. Am. Chem. Soc. 129 (26): 8058–8059.
  38. 38 Lefebvre, J., Austing, D.G., Bond, J. et al. (2006). Photoluminescence imaging of suspended single‐walled carbon nanotubes. Nano Lett. 6 (8): 1603–1608.
  39. 39 Lacerda, L., Pastorin, G., Wu, W. et al. (2006). Luminescence of functionalized carbon nanotubes as a tool to monitor bundle formation and dissociation in water: the effect of plasmid‐DNA complexation. Adv. Funct. Mater. 16 (14): 1839–1846.
  40. 40 Zhu, S., Song, Y., Wang, J. et al. (2017). Photoluminescence mechanism in graphene quantum dots: quantum confinement effect and surface/edge state. Nano Today 13: 10–14.
  41. 41 Dekaliuk, M.O., Viagin, O., Malyukin, Y.V. et al. (2014). Fluorescent carbon nanomaterials: “quantum dots” or nanoclusters? Phys. Chem. Chem. Phys. 16 (30): 16075–16084.
  42. 42 Cao, L., Meziani, M.J., Sahu, S. et al. (2013). Photoluminescence properties of graphene versus other carbon nanomaterials. Acc. Chem. Res. 46 (1): 171–180.
  43. 43 Havlik, J., Petrakova, V., Rehor, I. et al. (2013). Boosting nanodiamond fluorescence: towards development of brighter probes. Nanoscale 5 (8): 3208–3211.
  44. 44 Mochalin, V.N., Shenderova, O., Ho, D. et al. (2011). The properties and applications of nanodiamonds. Nat. Nanotechnol. 7 (1): 11–23.
  45. 45 Levi, N., Hantgan, R.R., Lively, M.O. et al. (2006). C60‐fullerenes: detection of intracellular photoluminescence and lack of cytotoxic effects. J. Nanobiotechnol. 4: 14.
  46. 46 Jeong, J., Jung, J., Choi, M. et al. (2012). Color‐tunable photoluminescent fullerene nanoparticles. Adv. Mater. 24 (15): 1999–2003.
  47. 47 Kwag, D.S., Park, K., Oh, T. et al. (2013). Hyaluronated fullerenes with photoluminescent and antitumoral activity. Chem. Commun. (Cambridge) 49 (3): 282–284.
  48. 48 Zheng, X.T., Ananthanarayanan, A., Luo, K.Q. et al. (2015). Glowing graphene quantum dots and carbon dots: properties, syntheses, and biological applications. Small 11 (14): 1620–1636.
  49. 49 Hola, K., Zhang, Y., Wang, Y. et al. (2014). Carbon dots—emerging light emitters for bioimaging, cancer therapy and optoelectronics. Nano Today 9 (5): 590–603.
  50. 50 Zhu, S., Zhang, J., Qiao, C. et al. (2011). Strongly green‐photoluminescent graphene quantum dots for bioimaging applications. Chem. Commun. (Cambridge) 47 (24): 6858–6860.
  51. 51 Cao, L., Wang, X., Meziani, M.J. et al. (2007). Carbon dots for multiphoton bioimaging. J. Am. Chem. Soc. 129 (37): 11318–11319.
  52. 52 Liu, H., Na, W., Liu, Z. et al. (2017). A novel turn‐on fluorescent strategy for sensing ascorbic acid using graphene quantum dots as fluorescent probe. Biosens. Bioelectron. 92: 229–233.
  53. 53 Fan, Z., Zhou, S., Garcia, C. et al. (2017). pH‐responsive fluorescent graphene quantum dots for fluorescence‐guided cancer surgery and diagnosis. Nanoscale 9 (15): 4928–4933.
  54. 54 Sun, H., Wu, L., Gao, N. et al. (2013). Improvement of photoluminescence of graphene quantum dots with a biocompatible photochemical reduction pathway and its bioimaging application. ACS Appl. Mater. Interfaces 5 (3): 1174–1179.
  55. 55 Lu, S., Sui, L., Liu, J. et al. (2017). Near‐infrared photoluminescent polymer‐carbon nanodots with two‐photon fluorescence. Adv. Mater. 29 (15).
  56. 56 Liu, Q., Guo, B., Rao, Z. et al. (2013). Strong two‐photon‐induced fluorescence from photostable, biocompatible nitrogen‐doped graphene quantum dots for cellular and deep‐tissue imaging. Nano Lett. 13 (6): 2436–2441.
  57. 57 Neves, V., Gerondopoulos, A., Heister, E. et al. (2012). Cellular localization, accumulation and trafficking of double‐walled carbon nanotubes in human prostate cancer cells. Nano Res. 5 (4): 223–234.
  58. 58 Leeuw, T.K., Reith, R.M., Simonette, R.A. et al. (2007). Single‐walled carbon nanotubes in the intact organism: near‐IR imaging and biocompatibility studies in Drosophila. Nano Lett. 7 (9): 2650–2654.
  59. 59 Welsher, K., Liu, Z., Sherlock, S.P. et al. (2009). A route to brightly fluorescent carbon nanotubes for near‐infrared imaging in mice. Nat. Nanotechnol. 4 (11): 773–780.
  60. 60 Sun, X., Liu, Z., Welsher, K. et al. (2008). Nano‐graphene oxide for cellular imaging and drug delivery. Nano Res. 1 (3): 203–212.
  61. 61 Fu, C.C., Lee, H.Y., Chen, K. et al. (2007). Characterization and application of single fluorescent nanodiamonds as cellular biomarkers. Proc. Natl. Acad. Sci. U. S. A. 104 (3): 727–732.
  62. 62 Yu, S.J., Kang, M.W., Chang, H.C. et al. (2005). Bright fluorescent nanodiamonds: no photobleaching and low cytotoxicity. J. Am. Chem. Soc. 127 (50): 17604–17605.
  63. 63 Faklaris, O., Joshi, V., Irinopoulou, T. et al. (2009). Photoluminescent diamond nanoparticles for cell labeling: study of the uptake mechanism in mammalian cells. ACS Nano 3 (12): 3955–3962.
  64. 64 Mohan, N., Chen, C.S., Hsieh, H.H. et al. (2010). In vivo imaging and toxicity assessments of fluorescent nanodiamonds in Caenorhabditis elegans. Nano Lett. 10 (9): 3692–3699.
  65. 65 Mohan, N., Zhang, B., Chang, C.‐C. et al. (2011). Fluorescent nanodiamond – a novel nanomaterial for in vivo applications. MRS Proc. 1362.
  66. 66 Vaijayanthimala, V., Cheng, P.Y., Yeh, S.H. et al. (2012). The long‐term stability and biocompatibility of fluorescent nanodiamond as an in vivo contrast agent. Biomaterials 33 (31): 7794–7802.
  67. 67 Pantarotto, D., Briand, J.P., Prato, M. et al. (2004). Translocation of bioactive peptides across cell membranes by carbon nanotubes. Chem. Commun. (Cambridge) (1): 16–17.
  68. 68 Nakayama‐Ratchford, N., Bangsaruntip, S., Sun, X. et al. (2007). Noncovalent functionalization of carbon nanotubes by fluorescein‐polyethylene glycol: supramolecular conjugates with pH‐dependent absorbance and fluorescence. J. Am. Chem. Soc. 129 (9): 2448–2449.
  69. 69 Kam, N.W., Liu, Z., and Dai, H. (2006). Carbon nanotubes as intracellular transporters for proteins and DNA: an investigation of the uptake mechanism and pathway. Angew. Chem. Int. Ed. 45 (4): 577–581.
  70. 70 Yang, K., Zhang, S., Zhang, G. et al. (2010). Graphene in mice: ultrahigh in vivo tumor uptake and efficient photothermal therapy. Nano Lett. 10 (9): 3318–3323.
  71. 71 Yang, K., Wan, J., Zhang, S. et al. (2011). In vivo pharmacokinetics, long‐term biodistribution, and toxicology of PEGylated graphene in mice. ACS Nano 5 (1): 516–522.
  72. 72 Hong, H., Yang, K., Zhang, Y. et al. (2012). In vivo targeting and imaging of tumor vasculature with radiolabeled, antibody‐conjugated nanographene. ACS Nano 6 (3): 2361–2370.
  73. 73 Chang, I.P., Hwang, K.C., and Chiang, C.S. (2008). Preparation of fluorescent magnetic nanodiamonds and cellular imaging. J. Am. Chem. Soc. 130 (46): 15476–15481.
  74. 74 Zhang, X.Q., Lam, R., Xu, X. et al. (2011). Multimodal nanodiamond drug delivery carriers for selective targeting, imaging, and enhanced chemotherapeutic efficacy. Adv. Mater. 23 (41): 4770–4775.
  75. 75 Salaam, A.D., Hwang, P., McIntosh, R. et al. (2014). Nanodiamond‐DGEA peptide conjugates for enhanced delivery of doxorubicin to prostate cancer. Beilstein J. Nanotechnol. 5: 937–945.
  76. 76 Rai, P., Mallidi, S., Zheng, X. et al. (2010). Development and applications of photo‐triggered theranostic agents. Adv. Drug Delivery Rev. 62 (11): 1094–1124.
  77. 77 Menon, J.U., Jadeja, P., Tambe, P. et al. (2013). Nanomaterials for photo‐based diagnostic and therapeutic applications. Theranostics 3 (3): 152–166.
  78. 78 Shen, Y., Shuhendler, A.J., Ye, D. et al. (2016). Two‐photon excitation nanoparticles for photodynamic therapy. Chem. Soc. Rev. 45 (24): 6725–6741.
  79. 79 Chakrabarti, M., Kiseleva, R., Vertegel, A. et al. (2015). Carbon nanomaterials for drug delivery and cancer therapy. J. Nanosci. Nanotechnol. 15 (8): 5501–5511.
  80. 80 Liu, Z., Chen, K., Davis, C. et al. (2008). Drug delivery with carbon nanotubes for in vivo cancer treatment. Cancer Res. 68 (16): 6652–6660.
  81. 81 Zou, L., Wang, H., He, B. et al. (2016). Current approaches of photothermal therapy in treating cancer metastasis with nanotherapeutics. Theranostics 6 (6): 762–772.
  82. 82 Chen, Q., Wen, J., Li, H. et al. (2016). Recent advances in different modal imaging‐guided photothermal therapy. Biomaterials 106: 144–166.
  83. 83 Chen, Y., Wang, L., and Shi, J. (2016). Two‐dimensional non‐carbonaceous materials‐enabled efficient photothermal cancer therapy. Nano Today 11 (3): 292–308.
  84. 84 Zheng, T., Li, G.G., Zhou, F. et al. (2016). Gold‐nanosponge‐based multistimuli‐responsive drug vehicles for targeted chemo‐photothermal therapy. Adv. Mater. 28 (37): 8218–8226.
  85. 85 Lee, C., Hwang, H.S., Lee, S. et al. (2017). Rabies virus‐inspired silica‐coated gold nanorods as a photothermal therapeutic platform for treating brain tumors. Adv. Mater. 29 (13): 1605563.
  86. 86 Chen, W., Zhang, S., Yu, Y. et al. (2016). Structural‐engineering rationales of gold nanoparticles for cancer theranostics. Adv. Mater. 28 (39): 8567–8585.
  87. 87 Ji, M., Xu, M., Zhang, W. et al. (2016). Structurally well‐defined Au@Cu2 − xS core‐shell nanocrystals for improved cancer treatment based on enhanced photothermal efficiency. Adv. Mater. 28 (16): 3094–3101.
  88. 88 Wang, Z., Huang, P., Jacobson, O. et al. (2016). Biomineralization‐inspired synthesis of copper sulfide‐ferritin nanocages as cancer theranostics. ACS Nano 10 (3): 3453–3460.
  89. 89 Riedinger, A., Avellini, T., Curcio, A. et al. (2015). Post‐synthesis incorporation of 64 Cu in CuS nanocrystals to radiolabel photothermal probes: a feasible approach for clinics. J. Am. Chem. Soc. 137 (48): 15145–15151.
  90. 90 Lyu, Y., Xie, C., Chechetka, S.A. et al. (2016). Semiconducting polymer nanobioconjugates for targeted photothermal activation of neurons. J. Am. Chem. Soc. 138 (29): 9049–9052.
  91. 91 Moon, H.K., Lee, S.H., and Choi, H.C. (2009). In vivo near‐infrared mediated tumor destruction by photothermal effect of carbon nanotubes. ACS Nano 3 (11): 3707–3713.
  92. 92 Robinson, J.T., Welsher, K., Tabakman, S.M. et al. (2010). High performance in vivo near‐IR (>1 mum) imaging and photothermal cancer therapy with carbon nanotubes. Nano Res. 3 (11): 779–793.
  93. 93 Wang, D., Hou, C., Meng, L. et al. (2017). Stepwise growth of gold coated cancer targeting carbon nanotubes for the precise delivery of doxorubicin combined with photothermal therapy. J. Mater. Chem. B 5 (7): 1380–1387.
  94. 94 Wang, S., Lin, Q., Chen, J. et al. (2017). Biocompatible polydopamine‐encapsulated gadolinium‐loaded carbon nanotubes for MRI and color mapping guided photothermal dissection of tumor metastasis. Carbon 112: 53–62.
  95. 95 Balandin, A.A. (2011). Thermal properties of graphene and nanostructured carbon materials. Nat. Mater. 10 (8): 569–581.
  96. 96 Robinson, J.T., Tabakman, S.M., Liang, Y. et al. (2011). Ultrasmall reduced graphene oxide with high near‐infrared absorbance for photothermal therapy. J. Am. Chem. Soc. 133 (17): 6825–6831.
  97. 97 Castano, A.P., Mroz, P., and Hamblin, M.R. (2006). Photodynamic therapy and anti‐tumour immunity. Nat. Rev. Cancer 6 (7): 535–545.
  98. 98 Celli, J.P., Spring, B.Q., Rizvi, I. et al. (2010). Imaging and photodynamic therapy: mechanisms, monitoring, and optimization. Chem. Rev. 110 (5): 2795–2838.
  99. 99 Detty, M.R., Gibson, S.L., and Wagner, S.J. (2004). Current clinical and preclinical photosensitizers for use in photodynamic therapy. J. Med. Chem. 47 (16): 3897–3915.
  100. 100 Markovic, Z. and Trajkovic, V. (2008). Biomedical potential of the reactive oxygen species generation and quenching by fullerenes (C60). Biomaterials 29 (26): 3561–3573.
  101. 101 Mroz, P., Xia, Y., Asanuma, D. et al. (2011). Intraperitoneal photodynamic therapy mediated by a fullerene in a mouse model of abdominal dissemination of colon adenocarcinoma. Nanomedicine 7 (6): 965–974.
  102. 102 Shi, J., Yu, X., Wang, L. et al. (2013). PEGylated fullerene/iron oxide nanocomposites for photodynamic therapy, targeted drug delivery and MR imaging. Biomaterials 34 (37): 9666–9677.
  103. 103 Murakami, T., Nakatsuji, H., Inada, M. et al. (2012). Photodynamic and photothermal effects of semiconducting and metallic‐enriched single‐walled carbon nanotubes. J. Am. Chem. Soc. 134 (43): 17862–17865.
  104. 104 Wang, L., Shi, J., Liu, R. et al. (2014). Photodynamic effect of functionalized single‐walled carbon nanotubes: a potential sensitizer for photodynamic therapy. Nanoscale 6 (9): 4642–4651.
  105. 105 Lee, D.J., Park, S.Y., Oh, Y.T. et al. (2011). Preparation of chlorine e6‐conjugated single‐wall carbon nanotube for photodynamic therapy. Macromol. Res. 19 (8): 848–852.
  106. 106 Tian, B., Wang, C., Zhang, S. et al. (2011). Photothermally enhanced photodynamic therapy delivered by nano‐graphene oxide. ACS Nano 5 (9): 7000–7009.
  107. 107 Huang, P., Xu, C., Lin, J. et al. (2011). Folic acid‐conjugated graphene oxide loaded with photosensitizers for targeting photodynamic therapy. Theranostics 1: 240–250.
  108. 108 Markovic, Z.M., Ristic, B.Z., Arsikin, K.M. et al. (2012). Graphene quantum dots as autophagy‐inducing photodynamic agents. Biomaterials 33 (29): 7084–7092.
  109. 109 Ristic, B.Z., Milenkovic, M.M., Dakic, I.R. et al. (2014). Photodynamic antibacterial effect of graphene quantum dots. Biomaterials 35 (15): 4428–4435.
  110. 110 Ge, J., Lan, M., Zhou, B. et al. (2014). A graphene quantum dot photodynamic therapy agent with high singlet oxygen generation. Nat. Commun. 5: 4596.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset