Next Article in Journal
A Gold Nanocage Probe Targeting Survivin for the Diagnosis of Pancreatic Cancer
Previous Article in Journal
Development of Biopredictive Dissolution Method for Extended-Release Desvenlafaxine Tablets
Previous Article in Special Issue
Orange-Peel-Derived Nanobiochar for Targeted Cancer Therapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Carbon Nanomaterials (CNMs) in Cancer Therapy: A Database of CNM-Based Nanocarrier Systems

School of Chemical Sciences, Dublin City University, Glasnevin, D09 NA55 Dublin, Ireland
*
Author to whom correspondence should be addressed.
Pharmaceutics 2023, 15(5), 1545; https://doi.org/10.3390/pharmaceutics15051545
Submission received: 17 April 2023 / Revised: 10 May 2023 / Accepted: 15 May 2023 / Published: 19 May 2023

Abstract

:
Carbon nanomaterials (CNMs) are an incredibly versatile class of materials that can be used as scaffolds to construct anticancer nanocarrier systems. The ease of chemical functionalisation, biocompatibility, and intrinsic therapeutic capabilities of many of these nanoparticles can be leveraged to design effective anticancer systems. This article is the first comprehensive review of CNM-based nanocarrier systems that incorporate approved chemotherapy drugs, and many different types of CNMs and chemotherapy agents are discussed. Almost 200 examples of these nanocarrier systems have been analysed and compiled into a database. The entries are organised by anticancer drug type, and the composition, drug loading/release metrics, and experimental results from these systems have been compiled. Our analysis reveals graphene, and particularly graphene oxide (GO), as the most frequently employed CNM, with carbon nanotubes and carbon dots following in popularity. Moreover, the database encompasses various chemotherapeutic agents, with antimicrotubule agents being the most common payload due to their compatibility with CNM surfaces. The benefits of the identified systems are discussed, and the factors affecting their efficacy are detailed.

1. Introduction

While cancer remains one of the world’s leading causes of death, advances in diagnostics and treatment have seen an overall improvement in detection and mortality rates. However, the current treatment approaches are either highly invasive, in the case of surgical operations, or can cause unwanted toxic side effects, as commonly experienced with chemotherapeutic agents and radiotherapy [1,2]. In particular, the effectiveness of chemotherapeutic agents is often limited by their poor aqueous solubility and nonselective nature, resulting in poor bioavailability and the indiscriminate death of both healthy and cancer cells [3]. To overcome these issues, there has been much research into the use of nanocarriers for the targeted and controlled release of anticancer drugs [4], where carbon nanomaterials (CNMs) have emerged in recent years as very promising candidates for this purpose. CNMs are a distinct class of materials that show altered characteristics to those of bulk carbon materials, such as diamond or graphite. They are classified as 0D, 1D, or 2D, according to the number of dimensions they possess which exist on the nanoscale (<100 nm) [5]. The allotropic nature of carbon means that a variety of these materials exists, some examples of which include graphene [1], carbon nanotubes (CNTs) [6], carbon nano-onions (CNOs) [7], nanodiamonds (NDs) [8], and carbon nanohorns [9]. CNMs have garnered widespread attention for their biomedical applications, such as drug delivery and diagnostics, because of their unique and highly desirable physicochemical and mechanical properties, such as size, biocompatibility, high tensile strength, and ease of chemical functionalisation.
By carefully selecting the production method, the particle sizes of CNMs can be precisely controlled, allowing for the creation of particles comparable in size to biomolecules (<100 nm) [10]. This size control enables CNM-based particles to take advantage of the leaky vasculature surrounding tumour cells through the enhanced permeation and retention (EPR) effect, facilitating the passive targeting of tumour cells [11]. However, passive targeting is generally limited, as not all tumours exhibit the EPR effect, and the random nature of the process makes it difficult to control and can lead to drug resistance [4]. Instead, it is preferable to actively target tumour tissues using targeting moieties that improve drug uptake through mechanisms such as receptor-mediated endocytosis [12]. The advanced surface chemistries provided by CNMs enable the attachment of various targeting ligands (such as folic acid (FA) [13]), imaging agents (such as BODIPY [14]), and anticancer drug molecules (such as cisplatin [15]), facilitating the creation of multifunctional nanocarriers. These nanocarriers can efficiently target, image, and deliver therapeutic agents directly to cancer cells, capitalizing on the unique properties of CNMs. Therefore, CNMs can be used as scaffolds to create theragnostic systems, combining imaging, detection, and treatment modalities in one tiny package to effectively diagnose and treat various illnesses [16].
As mentioned previously, there are many options for functionalising CNMs, with oxidation being one of the most straightforward approaches. This method introduces hydroxyl, carbonyl, and carboxyl groups to the surface of the nanomaterial, allowing for further functionalisation, and significantly increasing the material’s aqueous solubility in the process. Highly soluble CNMs can be utilised to increase the solubility of hydrophobic drugs, an approach taken by Cakmak and Eroglu, who employed graphene oxide (GO) to solubilise tamoxifen [16]. Facilitating the delivery of poorly soluble and/or poorly permeable drugs is a major benefit of the nanocarrier approach, as it does not require extensive modification of the drug molecule itself.
A range of other examples of covalent CNM modification exists, such as amidation, fluorination, and alkylation [17]. Covalent functionalisation methods have their drawbacks, mainly because this type of modification can damage the nanomaterial’s surface [7]. This surface damage can lead to a loss in the CNM’s unique electronic and physical properties, which may be essential to the nanocarrier’s effectiveness.
To circumnavigate issues associated with covalent modification, noncovalent functionalisation methods have been employed to attach components to the CNM. In this case, interactions, such as π–π stacking and hydrophobic interactions, are used to bind a molecule to the CNM surface [18].
A crucial aspect to consider when attaching drug molecules to a CNM is the drug release mechanism, which includes factors such as the release trigger and drug release profile, ensuring controlled and targeted delivery. Noncovalent attachment is particularly suited to reversible drug binding and can be utilised to design pH-responsive [19], redox-responsive [20], and NIR-responsive [21] drug delivery systems. Due to the acidity of tumour microenvironments, pH-sensitive systems are particularly relevant to cancer therapy [22]. This approach can be used to release the bound drug exclusively in the target tumour tissues, reducing unwanted side effects. Covalent strategies, such as drug attachment via hydrolysable ester bonds, have also been used for pH-responsive drug delivery, and often have the benefit of reduced drug leakage at neutral pH [23].
Herein, we present a database of nearly 200 CNM-based nanocarriers that have been utilised as drug delivery systems for clinically approved anticancer drugs. We curated this database through a comprehensive literature analysis, the details of which are provided in the following section. The entries are organised by drug type, and the composition, experimental results, drug loading and release metrics, and biological study models used are detailed. We also provide a critical analysis and discussion of the database and explore possible future research directions in the utilisation of CNM-based nanocarriers for anticancer drug delivery.

2. Methods and Metrics Used to Construct the Database

2.1. Preparation of the Database

This database is an in-depth overview of carbon nanomaterial (CNM)-based anticancer drug delivery systems. To construct the database, CAS SciFindern [24] was utilised as the data source. Combinations of keywords, such as “carbon nanomaterial”, “carbon nanotube”, “chemotherapy drug”, “anticancer drug”, and “doxorubicin”, were used to gather references, and the Boolean operators “AND” and “OR” were used to combine these search terms. Only English research articles that specifically focused on using CNMs to deliver clinically approved anticancer drugs were selected. The following information was extracted from each paper and entered into the database: (1) the anticancer drug used; (2) the composition of the nanocarrier system, including the CNM, and any targeting ligands, fluorophores, dispersants, etc., that were used; (3) the in vitro, in vivo, and ex vivo biological study models that the nanocarrier was tested on, including cell lines and animal breeds; (4) the drug loading and release metrics; these were taken only when explicitly given in the paper and were not calculated in this review; (5) the experimental results and observations, which were typically taken from the Conclusions section of each paper. The references were grouped based on the anticancer drug used, and the database was organised by sorting these drugs alphabetically.

2.2. Drug Loading and Release Metrics

The therapeutic efficacy of a nanocarrier system depends on its ability to absorb and release anticancer drugs; as such, quantitative metrics are needed to measure these systems. Such metrics are used to describe and compare the drug loading and release capabilities of different nanocarrier systems in the database.
The drug loading content (DLC) describes the amount of drug loaded onto the nanocarrier (Equation (1)). It is important to note that whilst most studies use the total mass of the nanocarrier (the CNM base, plus the drug, plus any other components), some studies just use the mass of the CNM itself [25], which leads to artificially higher DLC values.
DLC   wt % = mass   of   drug   bound   to   nanocarrier total   mass   of   nanocarrier × 100
The drug loading efficiency (DLE), sometimes called the encapsulation or entrapment efficiency, is a measure of the effectiveness of the drug loading process and not a quantitative measure of the drug content (Equation (2)).
DLE   wt % = mass   of   drug   bound   to   nanocarrier total   mass   of   drug   added × 100
The drug release efficiency (DRE) quantifies the cumulative release of a therapeutic agent from the nanocarrier (Equation (3)). This is the total amount of bound drug released throughout the experiment.
DRE   wt % = total   mass   of   drug   released mass   of   drug   bound   to   nanocarrier × 100

3. Database of Carbon-Nanomaterial-Based Cancer Therapeutics

Herein, we present a database of CNM-based nanocarrier systems that transport clinically approved anticancer drugs, seen in Table 1 The database includes the composition of the nanocarrier, the in vitro and in vivo biological models the system was tested on, the drug loading and release metrics, and a summary of the experimental results. The database is organized alphabetically by the anticancer drug used in the formulation; an index can be seen in Table 2.

4. Discussion

A total of 38 approved anticancer drugs were used in CNM-based nanocarriers in the literature, a breakdown of which can be seen in Figure 1. The anticancer drugs were further classified according to their mechanism of action, and the prevalence of each class is shown in Figure 1. The classes are as follows: (1) Alkylating agents—these work by adding alkyl groups to DNA, which can lead to DNA strand breaks and inhibit DNA replication; (2) Antimetabolites—these interfere with the synthesis of DNA, RNA, and proteins by mimicking essential cellular metabolites; (3) Natural products—this category consists of chemotherapeutic agents derived from natural sources, such as plants, microorganisms, or marine organisms. These agents often target specific aspects of cell division or DNA replication; (4) Hormone therapies—these target hormone-dependent cancers by interfering with the action of specific hormones or hormone receptors; (5) Antimicrotubule agents—these drugs target the microtubules, which play an essential role in cell division. By disrupting the formation or function of microtubules, these agents can inhibit cell division and lead to cancer cell death; (6) Miscellaneous agents—these include chemotherapeutic drugs that do not fit neatly into any of the other categories. The most diverse class of anticancer drugs used were alkylating agents, which is not surprising, as this class includes many Pt-based drugs, which can be easily complexed with oxidized CNMs. The miscellaneous-agent section contained five tyrosine kinase inhibitors, indicating the popularity of this class of drug. Hormone therapies were the least popular class of chemotherapy agents, with only three entries.
For each of the anticancer drugs in the database, several CNM-based nanocarriers have been investigated for their use in drug delivery. A total of 191 examples of CNM-based nanocarrier systems were found in the literature, many of which displayed higher anticancer efficacy with reduced side effects. As discussed in the introduction, the ease of functionalisation of CNM surfaces (graphene and CNT in particular) offers many different approaches to developing nanocarriers. A huge number of ligands were found to be used for drug delivery in the literature. These included polymers, such as PEG, which offer biocompatibility, water solubility, and reduced aggregation in situ [65,96], and biomolecules, such as folic acid, which enable the active targeting of folate receptors on tumour cells [177]. Other commonly used ligands were fluorescent agents, an example of which is Alexa Fluor, which is used for the fluorescent imaging of tumour cells [160], and peptides and proteins, offering improved bioavailability and stability [84]). Many of these approaches combined in the nanocarriers found in the literature show the complexity and breadth of options available to use.
The numerical analysis of the database shown in Figure 2 reveals that graphene (GO in particular) is the most popular class of CNMs to be incorporated into these systems, likely because it is one of the most well-established carbon nanomaterials. Over the years, a catalogue of functionalisation procedures has been developed, allowing for a range of moieties to be attached to the material’s surface [201]. The flat, aromatic surface of graphene lends itself excellently to π–π stacking, which allows for the easy noncovalent stacking of drug molecules. Graphene is also a PTT agent, a property that can be used to bolster the effectiveness of chemotherapy [14]. This is beneficial for the nanocarrier developed in [188], where free TAM is more efficacious than the nanocarrier-bound drug but the PTT potential of rGO makes it an attractive combinatorial therapy. Nonfunctionalised GO and rGO were found to successfully deliver anticancer drugs. GO and rGO showed pH- and NIR-triggered releases in some cases [32,132,148]. This offers a site-specific release of the drug, as the pH in tumour cells is typically lower than that in healthy cells, and NIR offers similar delivery.
CNTs came second in terms of popularity, which is surprising, as they are the oldest and most well-studied class of CNMs. This could be due to their tendency to aggregate into bundles in aqueous solutions, which could affect their biocompatibility. CNTs also do not have much in the way of intrinsic therapeutic or imaging properties; however, they do have large surface areas for drug loading. One example of pristine CNTs shows pH-dependent release and PTT [78]. Oxidised MWCNTs were found to be more toxic to healthy cells than cancer cells, which shows a need for further functionalisation [155]. The lack of control of intracellular accumulation also highlights the need for the attachment of targeting ligands to these systems.
Carbon dots, including GQDs, were the third most popular CNM to be used in these systems. The carbon allotropes that make up this new and rising class are tiny, even compared to other CNMs; this allows them to penetrate deep into cells to deliver drug molecules. They are even small enough to cross the blood–brain barrier [202], and their excellent biocompatibility [203] and ease of production in many cases [204] make them excellent nanocarrier scaffolds.
Many types of carbon dots are also water-soluble and have fluorescence imaging capabilities in the pristine form, for dual imaging and drug delivery [175]. Other types of CNMs, such as fullerenes, NDs, CNOs, HMCS, and CNHs, may not be as popular, but they show potential as nanocarriers due to their unique properties. For example, NDs offer improved biocompatibility over other CNMs [114,166], whilst fullerenes and CNOs display ease of functionalisation, narrow PDI, and ease of production [23,59].
In terms of drug moieties, antimicrotubule agents were the most popular payloads for these nanocarriers (Figure 2). This is due to the presence of multiple aromatic rings in these molecules, which facilitate noncovalent attachment to the CNM surface.
Anthracyclines, such as doxorubicin and epirubicin, are particularly popular. Alkylating agents are also quite popular, with platinum-based drugs often being complexed to the surface of the nanomaterial host. This is a particularly popular strategy with highly oxidised materials, such as GO, as the Pt can complex directly with oxygen-containing functional groups. Smaller hydrophilic organic molecules that lack any aromatic rings, such as those in the hormone therapy class, tend to not be so popular in CNM nanocarrier systems due to the lack of noncovalent interactions with the host CNM.
Designing systems that incorporate the intrinsic properties of CNMs allows for additional capabilities without having to chemically modify the material. The fluorescence of CDs and GQDs have proven to be useful for cellular imaging and tracking experiments [36,74,88]. Certain CNMs may also be utilised for killing cancer cells; for example, graphene has been used as a PTT agent, bolstering the effect of traditional chemotherapy [14,148]. Utilising this synergistic approach means that lower amounts of toxic chemotherapy drugs can be given to achieve the same therapeutic effect. The nπ* state of a CNM is essential for its intrinsic photothermal properties, and this state can be modulated by the addition of dopants (such as nitrogen) to the CNM [205]. Strong light absorption is required for a material to display photothermal properties, and a high photothermal conversion efficiency (η) is needed for a nanocarrier to be an effective PTT agent. For example, Forte et al. achieved photothermal-triggered drug release using a carbonised polymer dot-based system with an η value of 67.9% [206].
In general, systems that incorporate combination therapies exhibit some of the strongest anticancer effects due to the synergistic effects of PTT, PDT, single-drug and combination chemotherapy, or immunotherapy [67,129,169,190]. CNMs are the perfect class of nanoparticles for this approach, as they are easy to modify both covalently and noncovalently, with a range of functionalisation approaches available, allowing for the attachment of many different therapeutic agents. This, combined with the intrinsic properties of CNMs, can be leveraged to construct a range of nanocarrier systems.
The DLCs, DLEs, and DREs of nanocarriers are given as the number of entries above and below 50% in Figure 3. This was performed to make the entries more comparable given the differences in the sample sizes. A total of 152 out of 191 nanocarriers found in the literature had DLC, DLE, and/ or DRE data, and where DLC data were given, less than 35% of the nanocarriers were found to display values above 50%. This is similar for all CNMs, and surprisingly, graphene is the lowest, with only 22% of nanocarriers above 50% DLC. A much greater proportion of nanocarriers show DLEs above 50%, with fullerenes showing the lowest percentage of DLE values above 50%, and nanohorns displaying the highest at 100%; however, only one entry was available. This indicates that drug loading is an efficient process. In general, either a DLC or DLE value was given, with the DLE higher than the DLC.
The DRE, where given, was found to be quite high, on average, for all CNMs, with NDs displaying the lowest values; however, the sample size for this material was small compared to those for graphene and CNTs. Graphene-based nanocarriers incorporating Fe3O4 showed particularly high loading and release: [47,85,193]. GO-COOH also displayed high loading and release properties [133]. A lot of CNT-based systems with very high loading and release were observed; for example, the FA-PEG-bis-amine MWCNT system displayed 99% DLE and 90% DRE [27], SWCNTs showed 94% DLE and 93% DRE [78], and CNT-PEG displayed 95% DLC and 100% DRE [186].
For carbon dots, the CS-coated Fe3O4–NH2/GQD hybrid displayed 90% DLC + 84% DRE [33], whereas the GQD-HA system showed 98% DLE + 100% DRE [83], and the FA-CD-GOx nanocarrier yielded 82% DLE + 95% DRE [13]. Whilst there were much fewer fullerene systems with high drug loading/release metrics, acylated C60 fullerene displayed 81% DLC and 84% DRE [99]. A single ND system yielded 87% DLE and 80% DRE [166], whereas NHs and NPs do not show high combined loading and release.
The biocompatibility of these formulations must be further investigated, as certain CNMs are known to be toxic [207]. On the one hand, in the case of CNTs, the pristine form is toxic in mice and is dependent on the types of CNTs present [208]. On the other hand, pristine fullerenes such as C60 show no apparent toxicity, whereas some functionalised derivatives are highly toxic [209]. However, as previously discussed, the breadth of functionalisation methods and biocompatible ligands available to modify the surface chemistries of CNMs offers a variety of routes for overcoming this issue. The leakage of a drug at physiological pH is another issue that must be addressed in many systems, as toxic side effects are induced in vivo when the drug is released at neutral pH.

5. Conclusions and Future Directions

Overall, CNMs are incredibly versatile materials that can be used as both the foundation of nanocarrier systems and as therapeutic agents themselves. These systems can be designed to detect, image, and treat a range of tumours, from colorectal, brain, breast, liver, and stomach cancers. CNTs and graphene (GO in particular) were by far the most popular CNMs used in these systems due to their small size, high surface area, and ease of functionalisation. Other CNMs, such as carbon dots, are also growing in popularity due to their unique properties. A huge range of molecules, such as targeting ligands, fluorophores, dispersants, and drugs, can be easily attached to CNM surfaces, allowing for the construction of complex nanosystems.
Extensive in vivo biological work needs to be undertaken to fully understand the toxicity of these systems towards animals, and to overcome the regulatory hurdles needed to move these treatments into clinical trials.
The trend of designing theragnostic systems that incorporate the intrinsic properties of CNMs (such as PTT and fluorescence imaging) will likely be seen more in the future, as it allows for additional capabilities without damaging the CNM itself. Nanocarriers that leverage combinations of different therapies displayed the most potent anticancer effects, and therefore these systems will likely grow in popularity in the coming years.

Author Contributions

H.M. wrote the initial draft of the manuscript and prepared the figures and tables. A.F. assisted in proofreading and editing the document. S.G. supervised the project and assisted in writing and editing the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

The Irish Research Council (IRC) is gratefully acknowledged for providing a Government of Ireland Postgraduate Scholarship (GOIPG) (grant number GOIPG/2021/210) to H.M.

Acknowledgments

The authors wish to thank Michał Bartkowski for the fruitful discussions from the original idea, database planning, and final proofreading of the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

5-FU5-FluorouracilHPMCHydroxypropyl methylcellulose
6-MP6-MercaptopurineHSAHuman serum albumin
A1A549-cell-targeting oligonucleotideHYDHydrazone
ADAAdamantaneIMImatinib
ADHAdipic acid dihydrazideiRGDPEGylated RGD peptide
AFAlexa Fluor, AF488/647LALactobionic acid
ALAlendronateLELentinan
Anti-EpCAMEpithelial cell adhesion molecule antibodyLINLinoleic acid
APAAspartic acidMAMannose
AptAptamermAbAnti-VEGF monoclonal antibody
AQ4NBanoxantrone dihydrochloridemCNFMesoporous carbon nanoframe
AsoBcl-2 phosphorothioate antisense deoxyoligonucleotideMETMetformin
ATRAAll-trans retinoic acidMitPMitochondrion-targeting peptide
BSABovine serum albuminMMCMitomycin C
BTBiotinMMTMontmorillonite
CB[7]Cucurbit[7]urilMMT7Macrophage membrane hybridized with T7 peptide
CDCarbon dotMPAMercaptopropionic acid
CDTCyclodextrinMSCDMesoporous silica carbon dot
Ce6Chlorin e6MTXMitoxantrone
CisPCisplatinMETXMethotrexate
CMCCarboxymethyl celluloseMWCNTMultiwalled carbon nanotube
CNPCarbon nanoparticleNCCNanocrystalline cellulose
CNHCarbon nanohornNCNCNitrogen-doped carbon nanotube cup
CNMCarbon nanomaterialNDNanodiamond
CNRCarbon nanoringN-GONitrogen-doped graphene oxide
CNTCarbon nanotubeNGRAspargine–glycine–arginine peptide
COChito oligosaccharideNIRNear-infrared
CPCarboplatinNMCSNitrogen-doped mesoporous carbon sphere
CQDCarbon quantum dotN-prGONitrogen-doped porous reduced graphene oxide
cRGDCyclic RGD motifOPOxaliplatin
CSChitosanPAP-gp monoclonal antibody
CNSCarbon nanospherePAAPoly(acrylic acid)
CURCurcuminPAMAMPoly(amidoamine)
CWKG(KWKG)6H-(-CysTrp-Lys-Gly-)(-Lys-Trp-Lys-Gly-)6-OH peptidePANIPoly(aniline)
CβCDCarboxymethyl β-cyclodextrinPCPhosphatidylcholine
DCA-HPCHSDeoxycholic acid-modified hydroxypropyl chitosanPDAPolydopamine
DESDeep eutectic solventPDTPhotodynamic therapy
DLCDrug loading contentPEGPolyethylene glycol
DLEDrug loading efficiencyPEIPolyethyleneimine
DMPE1,2-dimyristoyl-sn-glycero-3-phosphoethanolaminePF68Pluronic F68
DOA2-(2-(docosyloxy)-2-oxoethoxy)acetic acidPGAPoly(glycolide)
DOXDoxorubicinP-gpP-glycoprotein antibodies
DPPTE1,2-dipalmitoyl-sn-glycero-3-phosphothioethanolPHEAEthanolamine
DREDrug release efficiencyPHEMAPolyhydroxyethyl methacrylate
DSPE1,2-distearoyl-sn-glycero-3-phosphoethanolaminePLAPoly(lactic acid)
DTXDocetaxelPLGAPoly(lactic-co-glycolic acid)
E2β-estradiolPMAAPoly(methacrylic acid)
EGFEpidermal growth factorPNMPoly(N-isopropylacrylamide)
EPIEpirubicinPNVCLPoly(N-vinylcaprolactam)
EtEtoposidePOEGMEAPoly(oligoethylene glycol methyl ether acrylate)
ExoCancer cell exosomesPRMPeptide protamine sulphate
FAFolic acidPTTPhotothermal therapy
FCDbBiotinylated Fe2+-doped carbon dotPTXPaclitaxel
f-CNTsFunctionalised carbon nanotubesPVAPoly(vinyl acetate)
FITCFluorescein isothiocyanatePVPPolyvinylpyrrolidone
FSCNOFucoidan-decorated silica–carbon nano-onionRBCRed blood cell membrane
GAGalactoseRfRiboflavin
GE11EGFR antagonist peptiderGOReduced graphene oxide
GEFGefitinibROSReactive oxygen species
GEMGemcitabineSASodium alginate
GFLGGly-phe-leu-gly enzyme-sensitive peptideSBSB-431542 (TGF-β inhibitor)
GGNGraphene–gold nanocompositesSLP2shRNA plasmid DNA
GlcNGlucosamineSWCNTSingle-walled carbon nanotube
GNRGold nanorodTAMTamoxifen
GOGraphene oxideTATTrans-activating transcriptional activator peptide
GOQDGraphene oxide quantum dotTAUTaurine
GOxGlucose oxidaseTEGTetra(ethylene glycol)
GQDGraphene quantum dotTfTransferrin
GRPGastrin-releasing peptideTMTemozolomide
HAHyaluronic acidTPTumour-targeting peptide (CKQFSALPFNFYT)
HIFAnti-hypoxia-inducible factor 1α antibodyTRTransferrin
HMHM30181A transmembrane P-glycoprotein inhibitorTTTopotecan
HMCSHollow mesoporous carbon sphereUAUrocanic acid
HPMCHydroxypropyl methylcelluloseVEGFVascular endothelial growth factor
HSAHuman serum albuminγPGAγ-polyglutamic acid

References

  1. Randviir, E.P.; Brownson, D.A.C.; Banks, C.E. A Decade of Graphene Research: Production, Applications and Outlook. Mater. Today 2014, 17, 426–432. [Google Scholar] [CrossRef]
  2. Seifalian, A. A New Era of Cancer Treatment: Carbon Nanotubes as Drug Delivery Tools. Int. J. Nanomed. 2011, 2011, 2963–2979. [Google Scholar] [CrossRef] [PubMed]
  3. Lukyanov, A.N.; Torchilin, V.P. Micelles from Lipid Derivatives of Water-Soluble Polymers as Delivery Systems for Poorly Soluble Drugs. Adv. Drug Deliv. Rev. 2004, 56, 1273–1289. [Google Scholar] [CrossRef] [PubMed]
  4. Peer, D.; Karp, J.M.; Hong, S.; Farokhzad, O.C.; Margalit, R.; Langer, R. Nanocarriers as an Emerging Platform for Cancer Therapy. Nat. Nanotech. 2007, 2, 751–760. [Google Scholar] [CrossRef] [PubMed]
  5. POTOČNIK, J. Commission Recommendation of 18 October 2011 on the Definition of Nanomaterial Text with EEA Relevance. Off. J. Eur. Union 2011, 275, 38–40. [Google Scholar]
  6. Yang, S.; Liu, J.; Ping, Y.; Wang, Z.; Zhang, J.; Zhang, L.; Cui, L.; Xiao, Y.; Qu, L. Multi-Functionalized Single-Walled Carbon Nanotubes as Delivery Carriers: Promote the Targeting Uptake and Antitumor Efficacy of Doxorubicin. J. Incl. Phenom. Macrocycl. Chem. 2022, 102, 801–817. [Google Scholar] [CrossRef]
  7. Mohan, H.; Bincoletto, V.; Arpicco, S.; Giordani, S. Supramolecular Functionalisation of B/N Co-Doped Carbon Nano-Onions for Novel Nanocarrier Systems. Materials 2022, 15, 5987. [Google Scholar] [CrossRef] [PubMed]
  8. Ye, W.; Han, H.; Li, H.; Jin, Q.; Wu, Y.; Chakrabortty, S.; Weil, T.; Ji, J. Polymer Coated Nanodiamonds as Gemcitabine Prodrug with Enzymatic Sensitivity for Pancreatic Cancer Treatment. Prog. Nat. Sci. Mater. Int. 2020, 30, 711–717. [Google Scholar] [CrossRef]
  9. Ma, X.; Shu, C.; Guo, J.; Pang, L.; Su, L.; Fu, D.; Zhong, W. Targeted Cancer Therapy Based on Single-Wall Carbon Nanohorns with Doxorubicin in Vitro and in Vivo. J. Nanopart. Res. 2014, 16, 2497. [Google Scholar] [CrossRef]
  10. Bartkowski, M.; Giordani, S. Carbon Nano-Onions as Potential Nanocarriers for Drug Delivery. Dalton Trans. 2021, 50, 2300–2309. [Google Scholar] [CrossRef]
  11. Shinde, V.R.; Revi, N.; Murugappan, S.; Singh, S.P.; Rengan, A.K. Enhanced Permeability and Retention Effect: A Key Facilitator for Solid Tumor Targeting by Nanoparticles. Photodiagn. Photodyn. Ther. 2022, 39, 102915. [Google Scholar] [CrossRef] [PubMed]
  12. Yu, Y.; Song, M.; Chen, C.; Du, Y.; Li, C.; Han, Y.; Yan, F.; Shi, Z.; Feng, S. Bortezomib-Encapsulated CuS/Carbon Dot Nanocomposites for Enhanced Photothermal Therapy via Stabilization of Polyubiquitinated Substrates in the Proteasomal Degradation Pathway. ACS Nano 2020, 14, 10688–10703. [Google Scholar] [CrossRef] [PubMed]
  13. Ghosh, A.; Ghosh, A.K.; Chowdhury, M.; Das, P.K. Folic Acid-Functionalized Carbon Dot-Enabled Starvation Therapy in Synergism with Paclitaxel against Breast Cancer. ACS Appl. Bio Mater. 2022, 5, 2389–2402. [Google Scholar] [CrossRef] [PubMed]
  14. Su, Y.; Wang, N.; Liu, B.; Du, Y.; Li, R.; Meng, Y.; Feng, Y.; Shan, Z.; Meng, S. A Phototheranostic Nanoparticle for Cancer Therapy Fabricated by BODIPY and Graphene to Realize Photo-Chemo Synergistic Therapy and Fluorescence/Photothermal Imaging. Dye Pigment. 2020, 177, 108262. [Google Scholar] [CrossRef]
  15. Fiekkies, J.T.R.; Fourie, E.; Erasmus, E. Cisplatin-Functionalized Nanodiamonds: Preparation and Characterization, with Potential Antineoplastic Application. Appl. Nanosci. 2021, 11, 2235–2245. [Google Scholar] [CrossRef]
  16. Keklikcioglu Cakmak, N.; Eroglu, A. Doxorubicin and Tamoxifen Loaded Graphene Oxide Nanoparticle Functionalized with Chitosan and Folic Acid for Anticancer Drug Delivery. Polym. Bull. 2023, 80, 2171–2185. [Google Scholar] [CrossRef]
  17. Bartelmess, J.; Giordani, S. Carbon Nano-Onions (Multi-Layer Fullerenes): Chemistry and Applications. Beilstein J. Nanotechnol. 2014, 5, 1980–1998. [Google Scholar] [CrossRef] [PubMed]
  18. Bartkowski, M.; Giordani, S. Supramolecular Chemistry of Carbon Nano-Onions. Nanoscale 2020, 12, 9352–9358. [Google Scholar] [CrossRef]
  19. Esmaeili, Y.; Seyedhosseini Ghaheh, H.; Ghasemi, F.; Shariati, L.; Rafienia, M.; Bidram, E.; Zarrabi, A. Graphene Oxide Quantum Dot-Chitosan Nanotheranostic Platform as a PH-Responsive Carrier for Improving Curcumin Uptake Internalization: In Vitro & in Silico Study. Biomater. Adv. 2022, 139, 213017. [Google Scholar] [CrossRef]
  20. Zhang, G.; Chang, H.; Amatore, C.; Chen, Y.; Jiang, H.; Wang, X. Apoptosis Induction and Inhibition of Drug Resistant Tumor Growth in Vivo Involving Daunorubicin-Loaded Graphene–Gold Composites. J. Mater. Chem. B 2013, 1, 493–499. [Google Scholar] [CrossRef]
  21. Gao, Q.; Gao, J.; Ding, C.; Li, S.; Deng, L.; Kong, Y. Construction of a PH- and near-Infrared Irradiation-Responsive Nanoplatform for Chemo-Photothermal Therapy. Int. J. Pharm. 2021, 593, 120112. [Google Scholar] [CrossRef] [PubMed]
  22. Arpicco, S.; Bartkowski, M.; Barge, A.; Zonari, D.; Serpe, L.; Milla, P.; Dosio, F.; Stella, B.; Giordani, S. Effects of the Molecular Weight of Hyaluronic Acid in a Carbon Nanotube Drug Delivery Conjugate. Front. Chem. 2020, 8, 578008. [Google Scholar] [CrossRef] [PubMed]
  23. Bahuguna, S.; Kumar, M.; Sharma, G.; Kumar, R.; Singh, B.; Raza, K. Fullerenol-Based Intracellular Delivery of Methotrexate: A Water-Soluble Nanoconjugate for Enhanced Cytotoxicity and Improved Pharmacokinetics. AAPS PharmSciTech 2018, 19, 1084–1092. [Google Scholar] [CrossRef] [PubMed]
  24. CAS SciFindern. Available online: https://scifinder-n.cas.org/ (accessed on 12 April 2023).
  25. Wang, R.; Cui, H.; Wang, J.; Li, N.; Zhao, Q.; Zhou, Y.; Lv, Z.; Zhong, W. Enhancing the Antitumor Effect of Methotrexate in Intro and in Vivo by a Novel Targeted Single-Walled Carbon Nanohorn-Based Drug Delivery System. RSC Adv. 2016, 6, 47272–47280. [Google Scholar] [CrossRef]
  26. Zavareh, H.S.; Pourmadadi, M.; Moradi, A.; Yazdian, F.; Omidi, M. Chitosan/Carbon Quantum Dot/Aptamer Complex as a Potential Anticancer Drug Delivery System towards the Release of 5-Fluorouracil. Int. J. Biol. Macromol. 2020, 165, 1422–1430. [Google Scholar] [CrossRef]
  27. Kaur, S.; Mehra, N.K.; Jain, K.; Jain, N.K. Development and Evaluation of Targeting Ligand-Anchored CNTs as Prospective Targeted Drug Delivery System. Artif. Cells Nanomed. Biotechnol. 2017, 45, 242–250. [Google Scholar] [CrossRef]
  28. Nivethaa, E.A.K.; Dhanavel, S.; Narayanan, V.; Narayana Kalkura, S.; Sivasankari, J.; Sivanandham, N.; Stephen, A. CS/Au/MWCNT Nanohybrid as an Efficient Carrier for the Sustained Release of 5-FU and a Study of Its Cytotoxicity on MCF-7. RSC Adv. 2021, 11, 4584–4592. [Google Scholar] [CrossRef]
  29. Garg, S.; Garg, A.; Sahu, N.K.; Yadav, A.K. Synthesis and Characterization of Nanodiamond-Anticancer Drug Conjugates for Tumor Targeting. Diam. Relat. Mater. 2019, 94, 172–185. [Google Scholar] [CrossRef]
  30. Wei, X.; Zhang, L.; Li, S.; Chen, X.; Zhang, M.; Wang, C.; Wang, T.; Li, L. A Designed Synthesis of Multifunctional Carbon Nanoframes for Simultaneous Imaging and Synergistic Chemo-Photothermal Cancer Therapy. New J. Chem. 2018, 42, 923–929. [Google Scholar] [CrossRef]
  31. Dou, Z.; Xu, Y.; Sun, H.; Liu, Y. Synthesis of PEGylated Fullerene–5-Fluorouracil Conjugates to Enhance the Antitumor Effect of 5-Fluorouracil. Nanoscale 2012, 4, 4624. [Google Scholar] [CrossRef]
  32. Fu, D.; You, J.; Guo, R.; Zhang, J.; Li, Q.; Wen, J.; Wang, H.; Yan, H. Preparation of Nanostructured Graphene Oxide and Its Application in Drug Loading and Sustained Release. ChemistrySelect 2022, 7, e202200670. [Google Scholar] [CrossRef]
  33. Hassani, S.; Gharehaghaji, N.; Divband, B. Chitosan-Coated Iron Oxide/Graphene Quantum Dots as a Potential Multifunctional Nanohybrid for Bimodal Magnetic Resonance/Fluorescence Imaging and 5-Fluorouracil Delivery. Mater. Today Commun. 2022, 31, 103589. [Google Scholar] [CrossRef]
  34. Hasanin, M.S.; El-Sakhawy, M.; Ahmed, H.Y.; Kamel, S. Hydroxypropyl Methylcellulose/Graphene Oxide Composite as Drug Carrier System for 5-fluorouracil. Biotechnol. J. 2022, 17, 2100183. [Google Scholar] [CrossRef] [PubMed]
  35. Pan, H.; Yu, Y.; Li, L.; Liu, B.; Liu, Y. Fabrication and Characterization of Taurine Functionalized Graphene Oxide with 5-Fluorouracil as Anticancer Drug Delivery Systems. Nanoscale Res. Lett. 2021, 16, 84. [Google Scholar] [CrossRef] [PubMed]
  36. Horo, H.; Saha, M.; Das, H.; Mandal, B.; Kundu, L.M. Synthesis of Highly Fluorescent, Amine-Functionalized Carbon Dots from Biotin-Modified Chitosan and Silk-Fibroin Blend for Target-Specific Delivery of Antitumor Agents. Carbohydr. Polym. 2022, 277, 118862. [Google Scholar] [CrossRef] [PubMed]
  37. Panda, S.; Bhol, C.S.; Bhutia, S.K.; Mohapatra, S. DSPE-PEG-Coated Uniform Nitrogen-Doped Carbon Capsules for NIR-Mediated Synergistic Chemophototherapy of Skin Cancer. ACS Appl. Bio Mater. 2021, 4, 7059–7069. [Google Scholar] [CrossRef]
  38. Talib, A.B.; Mohammed, M.H. Preparation, Characterization and Preliminary Cytotoxic Evaluation of 6-Mercaptopurine-Coated Biotinylated Carbon Dots Nanoparticles as a Drug Delivery System. Mater. Today Proc. 2021, 80, 2327–2333. [Google Scholar] [CrossRef]
  39. Chaudhari, N.S.; Pandey, A.P.; Patil, P.O.; Tekade, A.R.; Bari, S.B.; Deshmukh, P.K. Graphene Oxide Based Magnetic Nanocomposites for Efficient Treatment of Breast Cancer. Mater. Sci. Eng. C 2014, 37, 278–285. [Google Scholar] [CrossRef]
  40. González-Domínguez, J.M.; Grasa, L.; Frontiñán-Rubio, J.; Abás, E.; Domínguez-Alfaro, A.; Mesonero, J.E.; Criado, A.; Ansón-Casaos, A. Intrinsic and Selective Activity of Functionalized Carbon Nanotube/Nanocellulose Platforms against Colon Cancer Cells. Colloids Surf. B Biointerfaces 2022, 212, 112363. [Google Scholar] [CrossRef]
  41. Randive, D.S.; Gavade, A.S.; Shejawal, K.P.; Bhutkar, M.A.; Bhinge, S.D.; Jadhav, N.R. Colon Targeted Dosage Form of Capecitabine Using Folic Acid Anchored Modified Carbon Nanotube: In Vitro Cytotoxicity, Apoptosis and in Vivo Roentgenographic Study. Drug Dev. Ind. Pharm. 2021, 47, 1401–1412. [Google Scholar] [CrossRef]
  42. Makharza, S.; Cirillo, G.; Bachmatiuk, A.; Vittorio, O.; Mendes, R.G.; Oswald, S.; Hampel, S.; Rümmeli, M.H. Size-Dependent Nanographene Oxide as a Platform for Efficient Carboplatin Release. J. Mater. Chem. B 2013, 1, 6107. [Google Scholar] [CrossRef] [PubMed]
  43. Makharza, S.; Vittorio, O.; Cirillo, G.; Oswald, S.; Hinde, E.; Kavallaris, M.; Büchner, B.; Mertig, M.; Hampel, S. Graphene Oxide-Gelatin Nanohybrids as Functional Tools for Enhanced Carboplatin Activity in Neuroblastoma Cells. Pharm. Res. 2015, 32, 2132–2143. [Google Scholar] [CrossRef] [PubMed]
  44. Salas-Treviño, D.; Saucedo-Cárdenas, O.; Montes-de-Oca-Luna, R.; Rodríguez-Rocha, H.; García-García, A.; Montes-de-Oca-Luna, R.; Piña-Mendoza, E.I.; Contreras-Torres, F.F.; García-Rivas, G.; Soto-Domínguez, A. Hyaluronate Functionalized Multi-Wall Carbon Nanotubes Filled with Carboplatin as a Novel Drug Nanocarrier against Murine Lung Cancer Cells. Nanomaterials 2019, 9, 1572. [Google Scholar] [CrossRef] [PubMed]
  45. Balas, M.; Constanda, S.; Duma-Voiculet, A.; Prodana, M.; Hermenean, A.; Pop, S.; Demetrescu, I.; Dinischiotu, A. Fabrication and Toxicity Characterization of a Hybrid Material Based on Oxidized and Aminated MWCNT Loaded with Carboplatin. Toxicol. Vitr. 2016, 37, 189–200. [Google Scholar] [CrossRef]
  46. Liu, Y.; Pu, Y.; Sun, L.; Yao, H.; Zhao, B.; Zhang, R.; Zhang, Y. Folic Acid Functionalized γ Cyclodextrin C60, A Novel Vehicle for Tumor-Targeted Drug Delivery. J. Biomed. Nanotechnol. 2016, 12, 1393–1403. [Google Scholar] [CrossRef]
  47. Ramadan, I.; Nassar, M.Y.; Gomaa, A. In-Vitro Investigation of the Anticancer Efficacy of Carboplatin-Loaded Chitosan Nanocomposites Against Breast and Liver Cancer Cell Lines. J. Polym. Env. 2022, 31, 1102–1115. [Google Scholar] [CrossRef]
  48. Wei, L.; Li, G.; Lu, T.; Wei, Y.; Nong, Z.; Wei, M.; Pan, X.; Qin, Q.; Meng, F.; Li, X. Functionalized Graphene Oxide as Drug Delivery Systems for Platinum Anticancer Drugs. J. Pharm. Sci. 2021, 110, 3631–3638. [Google Scholar] [CrossRef]
  49. Yang, Y.-F.; Meng, F.-Y.; Li, X.-H.; Wu, N.-N.; Deng, Y.-H.; Wei, L.-Y.; Zeng, X.-P. Magnetic Graphene Oxide-Fe3O4 -PANI Nanoparticle Adsorbed Platinum Drugs as Drug Delivery Systems for Cancer Therapy. J. Nanosci. Nanotechnol. 2019, 19, 7517–7525. [Google Scholar] [CrossRef]
  50. Chen, J.-P.; Lu, Y.-J.; Hung, S.-C.; Chen, J.-P.; Wei, K.-C. Improving Thermal Stability and Efficacy of BCNU in Treating Glioma Cells Using PAA-Functionalized Graphene Oxide. Int. J. Nanomed. 2012, 2020, 1737–1747. [Google Scholar] [CrossRef]
  51. Salazar, A.; Pérez-de la Cruz, V.; Muñoz-Sandoval, E.; Chavarria, V.; García Morales, M.d.L.; Espinosa-Bonilla, A.; Sotelo, J.; Jiménez-Anguiano, A.; Pineda, B. Potential Use of Nitrogen-Doped Carbon Nanotube Sponges as Payload Carriers Against Malignant Glioma. Nanomaterials 2021, 11, 1244. [Google Scholar] [CrossRef]
  52. Singh, G.; Nenavathu, B.P.; Imtiyaz, K.; Moshahid A Rizvi, M. Fabrication of Chlorambucil Loaded Graphene- Oxide Nanocarrier and Its Application for Improved Antitumor Activity. Biomed. Pharmacother. 2020, 129, 110443. [Google Scholar] [CrossRef] [PubMed]
  53. Ezzati, N.; Mahjoub, A.R.; Shokrollahi, S.; Amiri, A.; Abolhosseini Shahrnoy, A. Novel Biocompatible Amino Acids-Functionalized Three-Dimensional Graphene Foams: As the Attractive and Promising Cisplatin Carriers for Sustained Release Goals. Int. J. Pharm. 2020, 589, 119857. [Google Scholar] [CrossRef] [PubMed]
  54. Ajima, K.; Yudasaka, M.; Murakami, T.; Maigné, A.; Shiba, K.; Iijima, S. Carbon Nanohorns as Anticancer Drug Carriers. Mol. Pharm. 2005, 2, 475–480. [Google Scholar] [CrossRef] [PubMed]
  55. Isaac, K.M.; Sabaraya, I.V.; Ghousifam, N.; Das, D.; Pekkanen, A.M.; Romanovicz, D.K.; Long, T.E.; Saleh, N.B.; Rylander, M.N. Functionalization of Single-Walled Carbon Nanohorns for Simultaneous Fluorescence Imaging and Cisplatin Delivery in Vitro. Carbon 2018, 138, 309–318. [Google Scholar] [CrossRef]
  56. Guven, A.; Villares, G.J.; Hilsenbeck, S.G.; Lewis, A.; Landua, J.D.; Dobrolecki, L.E.; Wilson, L.J.; Lewis, M.T. Carbon Nanotube Capsules Enhance the in Vivo Efficacy of Cisplatin. Acta Biomater. 2017, 58, 466–478. [Google Scholar] [CrossRef]
  57. Cui, Z.; Zhang, Y.; Zhang, J.; Kong, H.; Tang, X.; Pan, L.; Xia, K.; Aldalbahi, A.; Li, A.; Tai, R.; et al. Sodium Alginate-Functionalized Nanodiamonds as Sustained Chemotherapeutic Drug-Release Vectors. Carbon 2016, 97, 78–86. [Google Scholar] [CrossRef]
  58. Chen, X.; Li, D.; Wang, H.; Jiao, Y.; Wang, H.; Yu, Y.; Zhi, J. Fabrication of an EGF Modified Nanodiamonds-Based Anti-Cancer Drug Targeted Delivery System and Drug Carrier Uptake Visualization by 3D Raman Microscopy. RSC Adv. 2016, 6, 44543–44551. [Google Scholar] [CrossRef]
  59. Lynchak, O.; Byelinska, I.; Dziubenko, N.; Kuznietsova, H.; Abramchuk, O.; Prylutska, S. Acute Toxicity of C60–Cis-Pt Nanocomplex in Vivo. Appl. Nanosci. 2022, 12, 439–447. [Google Scholar] [CrossRef]
  60. Prylutska, S.; Grynyuk, I.; Skaterna, T.; Horak, I.; Grebinyk, A.; Drobot, L.; Matyshevska, O.; Senenko, A.; Prylutskyy, Y.; Naumovets, A.; et al. Toxicity of C60 Fullerene–Cisplatin Nanocomplex against Lewis Lung Carcinoma Cells. Arch. Toxicol. 2019, 93, 1213–1226. [Google Scholar] [CrossRef]
  61. Huang, H.; Liu, M.; Chen, J.; Mao, L.; Zeng, G.; Wen, Y.; Tian, J.; Zhou, N.; Zhang, X.; Wei, Y. Facile Fabrication of Carboxyl Groups Modified Fluorescent C 60 through a One-Step Thiol-Ene Click Reaction and Their Potential Applications for Biological Imaging and Intracellular Drug Delivery. J. Taiwan Inst. Chem. Eng. 2018, 86, 192–198. [Google Scholar] [CrossRef]
  62. Prylutska, S.; Panchuk, R.; Gołuński, G.; Skivka, L.; Prylutskyy, Y.; Hurmach, V.; Skorohyd, N.; Borowik, A.; Woziwodzka, A.; Piosik, J.; et al. C60 Fullerene Enhances Cisplatin Anticancer Activity and Overcomes Tumor Cell Drug Resistance. Nano Res. 2017, 10, 652–671. [Google Scholar] [CrossRef]
  63. Vasanthakumar, A.; Rejeeth, C.; Vivek, R.; Ponraj, T.; Jayaraman, K.; Anandasadagopan, S.K.; Vinayaga Moorthi, P. Design of Bio-Graphene-Based Multifunctional Nanocomposites Exhibits Intracellular Drug Delivery in Cervical Cancer Treatment. ACS Appl. Bio Mater. 2022, 5, 2956–2964. [Google Scholar] [CrossRef] [PubMed]
  64. Yu, C.; Long, Z.; Qiu, Q.; Liu, F.; Xu, Y.; Zhang, T.; Guo, R.; Zhong, W.; Huang, S.; Chen, S. Graphene Quantum Dots-Based Targeted Nanoprobes Detecting Drug Delivery, Imaging, and Enhanced Chemotherapy of Nasopharyngeal Carcinoma. Bioeng. Transl. Med. 2022, 7, e10270. [Google Scholar] [CrossRef]
  65. Giusto, E.; Žárská, L.; Beirne, D.F.; Rossi, A.; Bassi, G.; Ruffini, A.; Montesi, M.; Montagner, D.; Ranc, V.; Panseri, S. Graphene Oxide Nanoplatforms to Enhance Cisplatin-Based Drug Delivery in Anticancer Therapy. Nanomaterials 2022, 12, 2372. [Google Scholar] [CrossRef]
  66. Liu, P.; Xie, X.; Liu, M.; Hu, S.; Ding, J.; Zhou, W. A Smart MnO2-Doped Graphene Oxide Nanosheet for Enhanced Chemo-Photodynamic Combinatorial Therapy via Simultaneous Oxygenation and Glutathione Depletion. Acta Pharm. Sin. B 2021, 11, 823–834. [Google Scholar] [CrossRef] [PubMed]
  67. Astani, S.; Salehi, R.; Massoumi, B.; Massoudi, A. Co-Delivery of Cisplatin and Doxorubicin by Carboxylic Acid Functionalized Poly (Hydroxyethyl Methacrylate)/Reduced Graphene Nanocomposite for Combination Chemotherapy of Breast Cancer Cells. J. Biomater. Sci. Polym. Ed. 2021, 32, 657–677. [Google Scholar] [CrossRef]
  68. Zhang, Z.; Su, T.; Han, Y.; Yang, Z.; Wei, J.; Jin, L.; Fan, H. A Convergent Synthetic Platform for Dual Anticancer Drugs Functionalized by Reduced Graphene Nanocomposite Delivery for Hepatocellular Cancer. Drug Deliv. 2021, 28, 1982–1994. [Google Scholar] [CrossRef]
  69. Pei, X.; Zhu, Z.; Gan, Z.; Chen, J.; Zhang, X.; Cheng, X.; Wan, Q.; Wang, J. PEGylated Nano-Graphene Oxide as a Nanocarrier for Delivering Mixed Anticancer Drugs to Improve Anticancer Activity. Sci. Rep. 2020, 10, 2717. [Google Scholar] [CrossRef]
  70. Dučić, T.; Alves, C.S.; Vučinić, Ž.; Lázaro-Martínez, J.M.; Petković, M.; Soto, J.; Mutavdžić, D.; Valle Martínez de Yuso, M.; Radotić, K.; Algarra, M. S, N-Doped Carbon Dots-Based Cisplatin Delivery System in Adenocarcinoma Cells: Spectroscopical and Computational Approach. J. Colloid Interface Sci. 2022, 623, 226–237. [Google Scholar] [CrossRef]
  71. Zhou, C.; Li, H.; Liu, Y.; Wang, K. Design and Synthesis of Dual-Responsive Carbon Nanodots Loaded with Cisplatin for Targeted Therapy of Lung Cancer Therapy and Nursing Care. J. Clust. Sci. 2022, 33, 331–338. [Google Scholar] [CrossRef]
  72. Qi, Y.; Yang, W.; Liu, S.; Han, F.; Wang, H.; Zhao, Y.; Zhou, Y.; Zhou, D. Cisplatin Loaded Multiwalled Carbon Nanotubes Reverse Drug Resistance in NSCLC by Inhibiting EMT. Cancer Cell Int. 2021, 21, 74. [Google Scholar] [CrossRef] [PubMed]
  73. Li, Z.; Ni, J.; Liu, L.; Gu, L.; Wu, Z.; Li, T.; Ivanovich, K.I.; Zhao, W.; Sun, T.; Wang, T. Imaging-Guided Chemo–Photothermal Polydopamine Carbon Dots for EpCAM-Targeted Delivery toward Liver Tumor. ACS Appl. Mater. Interfaces 2021, 13, 29340–29348. [Google Scholar] [CrossRef] [PubMed]
  74. Zhang, B.; Duan, Q.; Li, Y.; Wang, J.; Zhang, W.; Sang, S. PH and Redox Dual-Sensitive Drug Delivery System Constructed Based on Fluorescent Carbon Dots. RSC Adv. 2021, 11, 2656–2663. [Google Scholar] [CrossRef] [PubMed]
  75. Fang, W.; Jin, R.; Mu, W. Near-Infrared Mediated Polymer-Coated Carbon Nanodots Loaded Cisplatin for Targeted Care Management of Lung Cancer Therapy. Process Biochem. 2020, 99, 27–35. [Google Scholar] [CrossRef]
  76. Razaghi, M.; Ramazani, A.; Khoobi, M.; Mortezazadeh, T.; Aksoy, E.A.; Küçükkılınç, T.T. Highly Fluorinated Graphene Oxide Nanosheets for Anticancer Linoleic-Curcumin Conjugate Delivery and T2-Weighted Magnetic Resonance Imaging: In Vitro and in Vivo Studies. J. Drug Deliv. Sci. Technol. 2020, 60, 101967. [Google Scholar] [CrossRef]
  77. Ghanbari, N.; Salehi, Z.; Khodadadi, A.A.; Shokrgozar, M.A.; Saboury, A.A. Glucosamine-Conjugated Graphene Quantum Dots as Versatile and PH-Sensitive Nanocarriers for Enhanced Delivery of Curcumin Targeting to Breast Cancer. Mater. Sci. Eng. C 2021, 121, 111809. [Google Scholar] [CrossRef]
  78. Li, H.; Zhang, N.; Hao, Y.; Wang, Y.; Jia, S.; Zhang, H.; Zhang, Y.; Zhang, Z. Formulation of Curcumin Delivery with Functionalized Single-Walled Carbon Nanotubes: Characteristics and Anticancer Effects in Vitro. Drug Deliv. 2014, 21, 379–387. [Google Scholar] [CrossRef]
  79. Li, H.; Zhang, N.; Hao, Y.; Wang, Y.; Jia, S.; Zhang, H. Enhancement of Curcumin Antitumor Efficacy and Further Photothermal Ablation of Tumor Growth by Single-Walled Carbon Nanotubes Delivery System in Vivo. Drug Deliv. 2019, 26, 1017–1026. [Google Scholar] [CrossRef]
  80. Consoli, G.M.L.; Giuffrida, M.L.; Satriano, C.; Musumeci, T.; Forte, G.; Petralia, S. A Novel Facile One-Pot Synthesis of Photothermally Responsive Carbon Polymer Dots as Promising Drug Nanocarriers. Chem. Commun. 2022, 58, 3126–3129. [Google Scholar] [CrossRef]
  81. Arvapalli, D.M.; Sheardy, A.T.; Allado, K.; Chevva, H.; Yin, Z.; Wei, J. Design of Curcumin Loaded Carbon Nanodots Delivery System: Enhanced Bioavailability, Release Kinetics, and Anticancer Activity. ACS Appl. Bio Mater. 2020, 3, 8776–8785. [Google Scholar] [CrossRef]
  82. Daneshmoghanlou, E.; Miralinaghi, M.; Moniri, E.; Sadjady, S.K. Fabrication of a PH-Responsive Magnetic Nanocarrier Based on Carboxymethyl Cellulose-Aminated Graphene Oxide for Loading and In-Vitro Release of Curcumin. J. Polym. Env. 2022, 30, 3718–3736. [Google Scholar] [CrossRef]
  83. Vahedi, N.; Tabandeh, F.; Mahmoudifard, M. Hyaluronic Acid–Graphene Quantum Dot Nanocomposite: Potential Target Drug Delivery and Cancer Cell Imaging. Biotech. App. Biochem. 2022, 69, 1068–1079. [Google Scholar] [CrossRef] [PubMed]
  84. Baneshi, M.; Dadfarnia, S.; Haji Shabani, A.M.; Sabbagh, S.K.; Bardania, H. AS1411 Aptamer-Functionalized Graphene Oxide-Based Nano-Carrier for Active-Target and PH-Sensitive Delivery of Curcumin. J. Iran. Chem. Soc. 2022, 19, 2367–2376. [Google Scholar] [CrossRef]
  85. Kazemi, S.; Pourmadadi, M.; Yazdian, F.; Ghadami, A. The Synthesis and Characterization of Targeted Delivery Curcumin Using Chitosan-Magnetite-Reduced Graphene Oxide as Nano-Carrier. Int. J. Biol. Macromol. 2021, 186, 554–562. [Google Scholar] [CrossRef]
  86. Gui, X.; Chen, Y.; Zhang, Z.; Lei, L.; Zhu, F.; Yang, W.; Guo, Y.; Chu, M. Fluorescent Hollow Mesoporous Carbon Spheres for Drug Loading and Tumor Treatment through 980-Nm Laser and Microwave Co-Irradiation. Biomaterials 2020, 248, 120009. [Google Scholar] [CrossRef]
  87. Azqhandi, M.H.A.; Farahani, B.V.; Dehghani, N. Encapsulation of Methotrexate and Cyclophosphamide in Interpolymer Complexes Formed between Poly Acrylic Acid and Poly Ethylene Glycol on Multi-Walled Carbon Nanotubes as Drug Delivery Systems. Mater. Sci. Eng. C 2017, 79, 841–847. [Google Scholar] [CrossRef]
  88. Sheng, Y.; Dai, W.; Gao, J.; Li, H.; Tan, W.; Wang, J.; Deng, L.; Kong, Y. PH-Sensitive Drug Delivery Based on Chitosan Wrapped Graphene Quantum Dots with Enhanced Fluorescent Stability. Mater. Sci. Eng. C 2020, 112, 110888. [Google Scholar] [CrossRef]
  89. Sima, L.E.; Chiritoiu, G.; Negut, I.; Grumezescu, V.; Orobeti, S.; Munteanu, C.V.A.; Sima, F.; Axente, E. Functionalized Graphene Oxide Thin Films for Anti-Tumor Drug Delivery to Melanoma Cells. Front. Chem. 2020, 8, 184. [Google Scholar] [CrossRef]
  90. Moore, T.L.; Grimes, S.W.; Lewis, R.L.; Alexis, F. Multilayered Polymer-Coated Carbon Nanotubes to Deliver Dasatinib. Mol. Pharm. 2014, 11, 276–282. [Google Scholar] [CrossRef]
  91. Hosseini, L.; Mahboobnia, K.; Irani, M. Fabrication of PLA/MWCNT/Fe3O4 Composite Nanofibers for Leukemia Cancer Cells. Int. J. Polym. Mater. Polym. Biomater. 2016, 65, 176–182. [Google Scholar] [CrossRef]
  92. Man, H.B.; Kim, H.; Kim, H.-J.; Robinson, E.; Liu, W.K.; Chow, E.K.-H.; Ho, D. Synthesis of Nanodiamond–Daunorubicin Conjugates to Overcome Multidrug Chemoresistance in Leukemia. Nanomed. Nanotechnol. Biol. Med. 2014, 10, 359–369. [Google Scholar] [CrossRef] [PubMed]
  93. Eskandari, S.; Barzegar, A.; Mahnam, K. Absorption of Daunorubicin and Etoposide Drugs by Hydroxylated and Carboxylated Carbon Nanotube for Drug Delivery: Theoretical and Experimental Studies. J. Biomol. Struct. Dyn. 2022, 40, 10057–10064. [Google Scholar] [CrossRef] [PubMed]
  94. Sheykhisarem, R.; Dehghani, H. In Vitro Biocompatibility Evaluations of PH-Sensitive Bi2MoO6/NH2-GO Conjugated Polyethylene Glycol for Release of Daunorubicin in Cancer Therapy. Colloids Surf. B Biointerfaces 2023, 221, 113006. [Google Scholar] [CrossRef] [PubMed]
  95. Lu, Y.; Wu, P.; Yin, Y.; Zhang, H.; Cai, C. Aptamer-Functionalized Graphene Oxide for Highly Efficient Loading and Cancer Cell-Specific Delivery of Antitumor Drug. J. Mater. Chem. B 2014, 2, 3849–3859. [Google Scholar] [CrossRef]
  96. Tas, A.; Keklikcioglu Cakmak, N. Synthesis of PEGylated Nanographene Oxide as a Nanocarrier for Docetaxel Drugs and Anticancer Activity on Prostate Cancer Cell Lines. Hum. Exp. Toxicol. 2021, 40, 172–182. [Google Scholar] [CrossRef]
  97. Li, B.; Zhang, X.-X.; Huang, H.-Y.; Chen, L.-Q.; Cui, J.-H.; Liu, Y.; Jin, H.; Lee, B.-J.; Cao, Q.-R. Effective Deactivation of A549 Tumor Cells in Vitro and in Vivo by RGD-Decorated Chitosan-Functionalized Single-Walled Carbon Nanotube Loading Docetaxel. Int. J. Pharm. 2018, 543, 8–20. [Google Scholar] [CrossRef]
  98. Zhao, Q.; Li, N.; Shu, C.; Li, R.; Ma, X.; Li, X.; Wang, R.; Zhong, W. Docetaxel-Loaded Single-Wall Carbon Nanohorns Using Anti-VEGF Antibody as a Targeting Agent: Characterization, in Vitro and in Vivo Antitumor Activity. J. Nanopart. Res. 2015, 17, 207. [Google Scholar] [CrossRef]
  99. Raza, K.; Thotakura, N.; Kumar, P.; Joshi, M.; Bhushan, S.; Bhatia, A.; Kumar, V.; Malik, R.; Sharma, G.; Guru, S.K.; et al. C60 -Fullerenes for Delivery of Docetaxel to Breast Cancer Cells: A Promising Approach for Enhanced Efficacy and Better Pharmacokinetic Profile. Int. J. Pharm. 2015, 495, 551–559. [Google Scholar] [CrossRef]
  100. Liu, W.; Ye, X.; He, L.; Cheng, J.; Luo, W.; Zheng, M.; Hu, Y.; Zhang, W.; Cao, Y.; Ran, H.; et al. A Novel Targeted Multifunctional Nanoplatform for Visual Chemo-Hyperthermia Synergy Therapy on Metastatic Lymph Nodes via Lymphatic Delivery. J. Nanobiotechnol. 2021, 19, 432. [Google Scholar] [CrossRef] [PubMed]
  101. Cheng, J.; Liu, Y.; He, L.; Liu, W.; Chen, Y.; Liu, F.; Guo, Y.; Ran, H.; Yang, L. Novel Multifunctional Nanoagent for Visual Chemo/Photothermal Therapy of Metastatic Lymph Nodes via Lymphatic Delivery. ACS Omega 2020, 5, 3194–3206. [Google Scholar] [CrossRef]
  102. Kim, D.; Kyung, H.; Lee, S.S.; Park, S.Y.; Yoo, W.C.; Lee, H.J. In Vitro Evaluation of Carbon-Based Nanospheres as Drug Delivery Vesicles in Breast Cancer Cell. Bull. Korean Chem. Soc. 2020, 41, 112–116. [Google Scholar] [CrossRef]
  103. Thotakura, N.; Sharma, S.; Khurana, R.K.; Babu, P.V.; Chitkara, D.; Kumar, V.; Singh, B.; Raza, K. Aspartic Acid Tagged Carbon Nanotubols as a Tool to Deliver Docetaxel to Breast Cancer Cells: Reduced Hemotoxicity with Improved Cytotoxicity. Toxicol. Vitr. 2019, 59, 126–134. [Google Scholar] [CrossRef] [PubMed]
  104. Sung, S.-Y.; Su, Y.-L.; Cheng, W.; Hu, P.-F.; Chiang, C.-S.; Chen, W.-T.; Hu, S.-H. Graphene Quantum Dots-Mediated Theranostic Penetrative Delivery of Drug and Photolytics in Deep Tumors by Targeted Biomimetic Nanosponges. Nano Lett. 2019, 19, 69–81. [Google Scholar] [CrossRef]
  105. Thotakura, N.; Sharma, G.; Singh, B.; Kumar, V.; Raza, K. Aspartic Acid Derivatized Hydroxylated Fullerenes as Drug Delivery Vehicles for Docetaxel: An Explorative Study. Artif. Cells Nanomed. Biotechnol. 2017, 46, 1763–1772. [Google Scholar] [CrossRef] [PubMed]
  106. Lv, Y.; Tao, L.; Annie Bligh, S.W.; Yang, H.; Pan, Q.; Zhu, L. Targeted Delivery and Controlled Release of Doxorubicin into Cancer Cells Using a Multifunctional Graphene Oxide. Mater. Sci. Eng. C Mater. Biol. Appl. 2016, 59, 652–660. [Google Scholar] [CrossRef] [PubMed]
  107. Xie, M.; Zhang, F.; Liu, L.; Zhang, Y.; Li, Y.; Li, H.; Xie, J. Surface Modification of Graphene Oxide Nanosheets by Protamine Sulfate/Sodium Alginate for Anti-Cancer Drug Delivery Application. Appl. Surf. Sci. 2018, 440, 853–860. [Google Scholar] [CrossRef]
  108. Lu, T.; Nong, Z.; Wei, L.; Wei, M.; Li, G.; Wu, N.; Liu, C.; Tang, B.; Qin, Q.; Li, X.; et al. Preparation and Anti-Cancer Activity of Transferrin/Folic Acid Double-Targeted Graphene Oxide Drug Delivery System. J. Biomater. Appl. 2020, 35, 15–27. [Google Scholar] [CrossRef]
  109. Qi, Z.; Shi, J.; Zhu, B.; Li, J.; Cao, S. Gold Nanorods/Graphene Oxide Nanosheets Immobilized by Polydopamine for Efficient Remotely Triggered Drug Delivery. J. Mater. Sci. 2020, 55, 14530–14543. [Google Scholar] [CrossRef]
  110. Chen, D.; Wang, C.; Jiang, F.; Liu, Z.; Shu, C.; Wan, L.-J. In Vitro and in Vivo Photothermally Enhanced Chemotherapy by Single-Walled Carbon Nanohorns as a Drug Delivery System. J. Mater. Chem. B 2014, 2, 4726–4732. [Google Scholar] [CrossRef]
  111. Liu, Z.; Fan, A.C.; Rakhra, K.; Sherlock, S.; Goodwin, A.; Chen, X.; Yang, Q.; Felsher, D.W.; Dai, H. Supramolecular Stacking of Doxorubicin on Carbon Nanotubes for In Vivo Cancer Therapy. Angew. Chem. Int. Ed. 2009, 48, 7668–7672. [Google Scholar] [CrossRef]
  112. Grebinyk, A.; Prylutska, S.; Grebinyk, S.; Prylutskyy, Y.; Ritter, U.; Matyshevska, O.; Dandekar, T.; Frohme, M. Complexation with C60 Fullerene Increases Doxorubicin Efficiency against Leukemic Cells In Vitro. Nanoscale Res. Lett. 2019, 14, 61. [Google Scholar] [CrossRef] [PubMed]
  113. Jović, D.S.; Seke, M.N.; Djordjevic, A.N.; Mrđanović, J.Ž.; Aleksić, L.D.; Bogdanović, G.M.; Pavić, A.B.; Plavec, J. Fullerenol Nanoparticles as a New Delivery System for Doxorubicin. RSC Adv. 2016, 6, 38563–38578. [Google Scholar] [CrossRef]
  114. Zhao, L.; Li, H.; Tan, L. A Novel Fullerene-Based Drug Delivery System Delivering Doxorubicin for Potential Lung Cancer Therapy. J. Nanosci. Nanotechnol. 2017, 17, 5147–5154. [Google Scholar] [CrossRef]
  115. Xiao, J.; Duan, X.; Yin, Q.; Zhang, Z.; Yu, H.; Li, Y. Nanodiamonds-Mediated Doxorubicin Nuclear Delivery to Inhibit Lung Metastasis of Breast Cancer. Biomaterials 2013, 34, 9648–9656. [Google Scholar] [CrossRef]
  116. Zhang, X.; Wang, S.; Fu, C.; Feng, L.; Ji, Y.; Tao, L.; Li, S.; Wei, Y. PolyPEGylated Nanodiamond for Intracellular Delivery of a Chemotherapeutic Drug. Polym. Chem. 2012, 3, 2716–2719. [Google Scholar] [CrossRef]
  117. Xu, Y.; Wang, C.; Wu, T.; Ran, G.; Song, Q. Template-Free Synthesis of Porous Fluorescent Carbon Nanomaterials with Gluten for Intracellular Imaging and Drug Delivery. ACS Appl. Mater. Interfaces 2022, 14, 21310–21318. [Google Scholar] [CrossRef]
  118. Thakur, C.K.; Neupane, R.; Karthikeyan, C.; Ashby, C.R.; Babu, R.J.; Boddu, S.H.S.; Tiwari, A.K.; Moorthy, N.S.H.N. Lysinated Multiwalled Carbon Nanotubes with Carbohydrate Ligands as an Effective Nanocarrier for Targeted Doxorubicin Delivery to Breast Cancer Cells. Molecules 2022, 27, 7461. [Google Scholar] [CrossRef] [PubMed]
  119. Wang, H.; Liang, Y.; Yin, Y.; Zhang, J.; Su, W.; White, A.M.; Jiang, B.; Xu, J.; Zhang, Y.; Stewart, S.; et al. Carbon Nano-Onion-Mediated Dual Targeting of P-Selectin and P-Glycoprotein to Overcome Cancer Drug Resistance. Nat. Commun. 2021, 12, 312. [Google Scholar] [CrossRef]
  120. Li, H.; Zeng, D.; Wang, Z.; Fang, L.; Li, F.; Wang, Z. Ultrasound-Enhanced Delivery of Doxorubicin/All-Trans Retinoic Acid-Loaded Nanodiamonds into Tumors. Nanomedicine 2018, 13, 981–996. [Google Scholar] [CrossRef]
  121. Du, X.; Li, L.; Wei, S.; Wang, S.; Li, Y. A Tumor-Targeted, Intracellular Activatable and Theranostic Nanodiamond Drug Platform for Strongly Enhanced in Vivo Antitumor Therapy. J. Mater. Chem. B 2020, 8, 1660–1671. [Google Scholar] [CrossRef]
  122. Falank, C.; Tasset, A.W.; Farrell, M.; Harris, S.; Everill, P.; Marinkovic, M.; Reagan, M.R. Development of Medical-Grade, Discrete, Multi-Walled Carbon Nanotubes as Drug Delivery Molecules to Enhance the Treatment of Hematological Malignancies. Nanomed. Nanotechnol. Biol. Med. 2019, 20, 102025. [Google Scholar] [CrossRef] [PubMed]
  123. Ganguly, S.; Das, P.; Itzhaki, E.; Hadad, E.; Gedanken, A.; Margel, S. Microwave-Synthesized Polysaccharide-Derived Carbon Dots as Therapeutic Cargoes and Toughening Agents for Elastomeric Gels. ACS Appl. Mater. Interfaces 2020, 12, 51940–51951. [Google Scholar] [CrossRef] [PubMed]
  124. Jiang, W.; Chen, J.; Gong, C.; Wang, Y.; Gao, Y.; Yuan, Y. Intravenous Delivery of Enzalutamide Based on High Drug Loading Multifunctional Graphene Oxide Nanoparticles for Castration-Resistant Prostate Cancer Therapy. J. Nanobiotechnol. 2020, 18, 50. [Google Scholar] [CrossRef] [PubMed]
  125. Yao, H.J.; Sun, L.; Liu, Y.; Jiang, S.; Pu, Y.; Li, J.; Zhang, Y. Monodistearoylphosphatidylethanolamine-Hyaluronic Acid Functionalization of Single-Walled Carbon Nanotubes for Targeting Intracellular Drug Delivery to Overcome Multidrug Resistance of Cancer Cells. Carbon 2016, 96, 362–376. [Google Scholar] [CrossRef]
  126. Suo, N.; Wang, M.; Jin, Y.; Ding, J.; Gao, X.; Sun, X.; Zhang, H.; Cui, M.; Zheng, J.; Li, N.; et al. Magnetic Multiwalled Carbon Nanotubes with Controlled Release of Epirubicin: An Intravesical Instillation System for Bladder Cancer. Int. J. Nanomed. 2019, 14, 1241–1254. [Google Scholar] [CrossRef]
  127. Wang, X.; Low, X.C.; Hou, W.; Abdullah, L.N.; Toh, T.B.; Mohd Abdul Rashid, M.; Ho, D.; Chow, E.K.-H. Epirubicin-Adsorbed Nanodiamonds Kill Chemoresistant Hepatic Cancer Stem Cells. ACS Nano 2014, 8, 12151–12166. [Google Scholar] [CrossRef]
  128. Hettiarachchi, S.D.; Graham, R.M.; Mintz, K.J.; Zhou, Y.; Vanni, S.; Peng, Z.; Leblanc, R.M. Triple Conjugated Carbon Dots as a Nano-Drug Delivery Model for Glioblastoma Brain Tumors. Nanoscale 2019, 11, 6192–6205. [Google Scholar] [CrossRef]
  129. Zhang, Y.-F.; Wu, Y.-F.; Lan, T.-J.; Chen, Y.; Su, S.-H. Codelivery of Anticancer Drug and Photosensitizer by PEGylated Graphene Oxide and Cell Penetrating Peptide Enhanced Tumor-Suppressing Effect on Osteosarcoma. Front. Mol. Biosci. 2021, 7, 618896. [Google Scholar] [CrossRef]
  130. Lan, M.-Y.; Hsu, Y.-B.; Lan, M.-C.; Chen, J.-P.; Lu, Y.-J. Polyethylene Glycol-Coated Graphene Oxide Loaded with Erlotinib as an Effective Therapeutic Agent for Treating Nasopharyngeal Cancer Cells. Int. J. Nanomed. 2020, 15, 7569–7582. [Google Scholar] [CrossRef]
  131. Lam, A.T.N.; Yoon, J.-H.; Ly, N.H.; Joo, S.-W. Electrostatically Self-Assembled Quinazoline-Based Anticancer Drugs on Negatively-Charged Nanodiamonds for Overcoming the Chemoresistances in Lung Cancer Cells. BioChip J. 2018, 12, 163–171. [Google Scholar] [CrossRef]
  132. Li, L.; Pan, C.; Guo, Z.; Liu, B.; Pan, H.; Wu, H.; Liu, Y. Exploration on the Adsorption Mechanism of Aminopyrimidines and Quinazoline Compounds on Graphene Oxide: Hydrophobicity and Structure-Controlled Release Process. Mat. Express 2019, 9, 419–428. [Google Scholar] [CrossRef]
  133. Gholami, A.; Emadi, F.; Nazem, M.; Aghayi, R.; Khalvati, B.; Amini, A.; Ghasemi, Y. Expression of Key Apoptotic Genes in Hepatocellular Carcinoma Cell Line Treated with Etoposide-Loaded Graphene Oxide. J. Drug Deliv. Sci. Technol. 2020, 57, 101725. [Google Scholar] [CrossRef]
  134. Wang, J.; Wang, R.; Zhang, F.; Yin, Y.; Mei, L.; Song, F.; Tao, M.; Yue, W.; Zhong, W. Overcoming Multidrug Resistance by a Combination of Chemotherapy and Photothermal Therapy Mediated by Carbon Nanohorns. J. Mater. Chem. B 2016, 4, 6043–6051. [Google Scholar] [CrossRef] [PubMed]
  135. Shirani, M.P.; Rezaei, B.; Khayamian, T.; Dinari, M.; Shamili, F.H.; Ramezani, M.; Alibolandi, M. Ingenious PH-Sensitive Etoposide Loaded Folic Acid Decorated Mesoporous Silica-Carbon Dot with Carboxymethyl-Βcyclodextrin Gatekeeper for Targeted Drug Delivery and Imaging. Mater. Sci. Eng. C 2018, 92, 892–901. [Google Scholar] [CrossRef]
  136. Heger, Z.; Polanska, H.; Krizkova, S.; Balvan, J.; Raudenska, M.; Dostalova, S.; Moulick, A.; Masarik, M.; Adam, V. Co-Delivery of VP-16 and Bcl-2-Targeted Antisense on PEG-Grafted OMWCNTs for Synergistic in Vitro Anti-Cancer Effects in Non-Small and Small Cell Lung Cancer. Colloids Surf. B Biointerfaces 2017, 150, 131–140. [Google Scholar] [CrossRef]
  137. Tiwari, H.; Karki, N.; Pal, M.; Basak, S.; Verma, R.K.; Bal, R.; Kandpal, N.D.; Bisht, G.; Sahoo, N.G. Functionalized Graphene Oxide as a Nanocarrier for Dual Drug Delivery Applications: The Synergistic Effect of Quercetin and Gefitinib against Ovarian Cancer Cells. Colloids Surf. B Biointerfaces 2019, 178, 452–459. [Google Scholar] [CrossRef]
  138. Gautam, A.; Pal, K. Gefitinib Conjugated PEG Passivated Graphene Quantum Dots Incorporated PLA Microspheres for Targeted Anticancer Drug Delivery. Heliyon 2022, 8, e12512. [Google Scholar] [CrossRef]
  139. More, M.P.; Deshmukh, P.K. Quality by Design Approach for the Synthesis of Graphene Oxide Nanosheets Using Full Factorial Design with Enhanced Delivery of Gefitinib Nanocrystals. Mater. Res. Express 2021, 8, 075602. [Google Scholar] [CrossRef]
  140. Liu, J.; Zhang, D.; Lian, S.; Zheng, J.; Li, B.; Li, T.; Jia, L. Redox-Responsive Hyaluronic Acid-Functionalized Graphene Oxide Nanosheets for Targeted Delivery of Water-Insoluble Cancer Drugs. Int. J. Nanomed. 2018, 13, 7457–7472. [Google Scholar] [CrossRef]
  141. Prajapati, S.K.; Jain, A.; Shrivastava, C.; Jain, A.K. Hyaluronic Acid Conjugated Multi-Walled Carbon Nanotubes for Colon Cancer Targeting. Int. J. Biol. Macromol. 2019, 123, 691–703. [Google Scholar] [CrossRef]
  142. Razzazan, A.; Atyabi, F.; Kazemi, B.; Dinarvand, R. In Vivo Drug Delivery of Gemcitabine with PEGylated Single-Walled Carbon Nanotubes. Mater. Sci. Eng. C 2016, 62, 614–625. [Google Scholar] [CrossRef] [PubMed]
  143. Pourjavadi, A.; Tehrani, Z.M.; Moghanaki, A.A. Folate-Conjugated PH-Responsive Nanocarrier Designed for Active Tumor Targeting and Controlled Release of Gemcitabine. Pharm. Res. 2016, 33, 417–432. [Google Scholar] [CrossRef] [PubMed]
  144. Lu, M.; Wang, Y.-K.; Zhao, J.; Lu, H.; Stenzel, M.H.; Xiao, P. PEG Grafted-Nanodiamonds for the Delivery of Gemcitabine. Macromol. Rapid Commun. 2016, 37, 2023–2029. [Google Scholar] [CrossRef] [PubMed]
  145. Lai, H.; Lu, M.; Lu, H.; Stenzel, M.H.; Xiao, P. PH-Triggered Release of Gemcitabine from Polymer Coated Nanodiamonds Fabricated by RAFT Polymerization and Copper Free Click Chemistry. Polym. Chem. 2016, 7, 6220–6230. [Google Scholar] [CrossRef]
  146. Farshi Azhar, F.; Rezaei, M.; Olad, A.; Mousazadeh, H. The Effect of Montmorillonite in Graphene Oxide/Chitosan Nanocomposite on Controlled Release of Gemcitabine. Polym. Bull. 2022, 79, 5861–5883. [Google Scholar] [CrossRef]
  147. Ashuri, A.; Miralinaghi, M.; Moniri, E. Evaluation of Folic Acid-Conjugated Chitosan Grafted Fe3O4/Graphene Oxide as a PH- and Magnetic Field-Responsive System for Adsorption and Controlled Release of Gemcitabine. Korean J. Chem. Eng. 2022, 39, 1880–1890. [Google Scholar] [CrossRef]
  148. Wei, X.; Li, P.; Zhou, H.; Hu, X.; Liu, D.; Wu, J.; Wang, Y. Engineering of Gemcitabine Coated Nano-Graphene Oxide Sheets for Efficient near-Infrared Radiation Mediated in Vivo Lung Cancer Photothermal Therapy. J. Photochem. Photobiol. B Biol. 2021, 216, 112125. [Google Scholar] [CrossRef]
  149. Yunus, U.; Zulfiqar, M.A.; Ajmal, M.; Bhatti, M.H.; Chaudhry, G.-S.; Muhammad, T.S.T.; Sung, Y.Y. Targeted Drug Delivery Systems: Synthesis and in Vitro Bioactivity and Apoptosis Studies of Gemcitabine-Carbon Dot Conjugates. Biomed. Mater. 2020, 15, 065004. [Google Scholar] [CrossRef]
  150. Zhang, P.; Yi, W.; Hou, J.; Yoo, S.; Jin, W.; Yang, Q. A Carbon Nanotube-Gemcitabine-Lentinan Three-Component Composite for Chemo-Photothermal Synergistic Therapy of Cancer. Int. J. Nanomed. 2018, 13, 3069–3080. [Google Scholar] [CrossRef]
  151. Tehrani, N.S.; Masoumi, M.; Chekin, F.; Baei, M.S. Nitrogen Doped Porous Reduced Graphene Oxide Hybrid as a Nanocarrier of Imatinib Anticancer Drug. Russ. J. Appl. Chem. 2020, 93, 1221–1228. [Google Scholar] [CrossRef]
  152. Yousefi, A.-M.; Safaroghli-Azar, A.; Fakhroueian, Z.; Bashash, D. ZnO/CNT@Fe3O4 Induces ROS-Mediated Apoptosis in Chronic Myeloid Leukemia (CML) Cells: An Emerging Prospective for Nanoparticles in Leukemia Treatment. Artif. Cells Nanomed. Biotechnol. 2020, 48, 735–745. [Google Scholar] [CrossRef]
  153. Felix, D.M.; Rebelo Alencar, L.M.; Duarte de Menezes, F.; Midlej, V.d.V.P.; Aguiar, L.; Gemini Piperni, S.; Zhang, J.; Liu, Y.; Ricci-Junior, E.; Alexis, F.; et al. Graphene Quantum Dots Decorated with Imatinib for Leukemia Treatment. J. Drug Deliv. Sci. Technol. 2021, 61, 102117. [Google Scholar] [CrossRef] [PubMed]
  154. Huang, Y.-S.; Lu, Y.-J.; Chen, J.-P. Magnetic Graphene Oxide as a Carrier for Targeted Delivery of Chemotherapy Drugs in Cancer Therapy. J. Magn. Magn. Mater. 2017, 427, 34–40. [Google Scholar] [CrossRef]
  155. Nicosia, A.; Cavallaro, G.; Costa, S.; Utzeri, M.; Cuttitta, A.; Giammona, G.; Mauro, N. Carbon Nanodots for On Demand Chemophotothermal Therapy Combination to Elicit Necroptosis: Overcoming Apoptosis Resistance in Breast Cancer Cell Lines. Cancers 2020, 12, 3114. [Google Scholar] [CrossRef] [PubMed]
  156. Chuang, C.-C.; Lan, Y.-H.; Lu, Y.-J.; Weng, Y.-L.; Chen, J.-P. Targeted Delivery of Irinotecan and SLP2 ShRNA with GRP-Conjugated Magnetic Graphene Oxide for Glioblastoma Treatment. Biomater. Sci. 2022, 10, 3201–3222. [Google Scholar] [CrossRef]
  157. Yu, S.; Zhang, Y.; Chen, L.; Li, Q.; Du, J.; Gao, Y.; Zhang, L.; Yang, Y. Antitumor Effects of Carbon Nanotube-drug Complex against Human Breast Cancer Cells. Exp. Ther. Med. 2018, 16, 1103–1110. [Google Scholar] [CrossRef]
  158. Yu, S.; Li, Q.; Wang, J.; Du, J.; Gao, Y.; Zhang, L.; Chen, L.; Yang, Y.; Liu, X. A Targeted Drug Delivery System Based on Carbon Nanotubes Loaded with Lobaplatin toward Liver Cancer Cells. J. Mater. Res. 2018, 33, 2565–2575. [Google Scholar] [CrossRef]
  159. Jiao, Z.; Zhang, B.; Li, C.; Kuang, W.; Zhang, J.; Xiong, Y.; Tan, S.; Cai, X.; Huang, L. Carboxymethyl Cellulose-Grafted Graphene Oxide for Efficient Antitumor Drug Delivery. Nanotechnol. Rev. 2018, 7, 291–301. [Google Scholar] [CrossRef]
  160. Das, M.; Datir, S.R.; Singh, R.P.; Jain, S. Augmented Anticancer Activity of a Targeted, Intracellularly Activatable, Theranostic Nanomedicine Based on Fluorescent and Radiolabeled, Methotrexate-Folic Acid-Multiwalled Carbon Nanotube Conjugate. Mol. Pharm. 2013, 10, 2543–2557. [Google Scholar] [CrossRef]
  161. Shan, S.; Jia, S.; Lawson, T.; Yan, L.; Lin, M.; Liu, Y. The Use of TAT Peptide-Functionalized Graphene as a Highly Nuclear-Targeting Carrier System for Suppression of Choroidal Melanoma. Int. J. Mol. Sci. 2019, 20, 4454. [Google Scholar] [CrossRef]
  162. D’souza, S.L.; Chettiar, S.S.; Koduru, J.R.; Kailasa, S.K. Synthesis of Fluorescent Carbon Dots Using Daucus Carota Subsp. Sativus Roots for Mitomycin Drug Delivery. Optik 2018, 158, 893–900. [Google Scholar] [CrossRef]
  163. Ohta, T.; Hashida, Y.; Yamashita, F.; Hashida, M. Sustained Release of Mitomycin C from Its Conjugate with Single-Walled Carbon Nanotubes Associated by Pegylated Peptide. Biol. Pharm. Bull. 2016, 39, 1687–1693. [Google Scholar] [CrossRef] [PubMed]
  164. Zhu, H.; Zhang, B.; Zhu, N.; Li, M.; Yu, Q. Mitochondrion Targeting Peptide-Modified Magnetic Graphene Oxide Delivering Mitoxantrone for Impairment of Tumor Mitochondrial Functions. Chin. Chem. Lett. 2021, 32, 1220–1223. [Google Scholar] [CrossRef]
  165. Risi, G.; Bloise, N.; Merli, D.; Icaro-Cornaglia, A.; Profumo, A.; Fagnoni, M.; Quartarone, E.; Imbriani, M.; Visai, L. Invitro Study of Multiwall Carbon Nanotubes (MWCNTs) with Adsorbed Mitoxantrone (MTO) as a Drug Delivery System to Treat Breast Cancer. RSC Adv. 2014, 4, 18683–18693. [Google Scholar] [CrossRef]
  166. Toh, T.-B.; Lee, D.-K.; Hou, W.; Abdullah, L.N.; Nguyen, J.; Ho, D.; Chow, E.K.-H. Nanodiamond–Mitoxantrone Complexes Enhance Drug Retention in Chemoresistant Breast Cancer Cells. Mol. Pharm. 2014, 11, 2683–2691. [Google Scholar] [CrossRef] [PubMed]
  167. Heister, E.; Neves, V.; Lamprecht, C.; Silva, S.R.P.; Coley, H.M.; McFadden, J. Drug Loading, Dispersion Stability, and Therapeutic Efficacy in Targeted Drug Delivery with Carbon Nanotubes. Carbon 2012, 50, 622–632. [Google Scholar] [CrossRef]
  168. Chen, Q.; Che, C.; Liu, J.; Gong, Z.; Si, M.; Yang, S.; Yang, G. Construction of an Exosome-Functionalized Graphene Oxide Based Composite Bionic Smart Drug Delivery System and Its Anticancer Activity. Nanotechnology 2022, 33, 175101. [Google Scholar] [CrossRef]
  169. Zhou, F.; Wang, M.; Luo, T.; Qu, J.; Chen, W.R. Photo-Activated Chemo-Immunotherapy for Metastatic Cancer Using a Synergistic Graphene Nanosystem. Biomaterials 2021, 265, 120421. [Google Scholar] [CrossRef]
  170. Kazempour, M.; Edjlali, L.; Akbarzadeh, A.; Davaran, S.; Farid, S.S. Synthesis and Characterization of Dual PH-and Thermo-Responsive Graphene-Based Nanocarrier for Effective Anticancer Drug Delivery. J. Drug Deliv. Sci. Technol. 2019, 54, 101158. [Google Scholar] [CrossRef]
  171. Farnaz, R.; Maryam, S.; Masoumeh, J.; Parvaneh, S. Colloidal HSA—Graphene Oxide Nanosheets for Sustained Release of Oxaliplatin: Preparation, Release Mechanism, Cytotoxicity and Electrochemical Approaches. Colloids Surf. B Biointerfaces 2018, 171, 10–16. [Google Scholar] [CrossRef]
  172. Wu, L.; Man, C.; Wang, H.; Lu, X.; Ma, Q.; Cai, Y.; Ma, W. PEGylated Multi-Walled Carbon Nanotubes for Encapsulation and Sustained Release of Oxaliplatin. Pharm. Res. 2013, 30, 412–423. [Google Scholar] [CrossRef] [PubMed]
  173. Ding, Y.-F.; Kwong, C.H.T.; Li, S.; Pan, Y.-T.; Wei, J.; Wang, L.-H.; Mok, G.S.P.; Wang, R. Supramolecular Nanomedicine Derived from Cucurbit[7]Uril-Conjugated Nano-Graphene Oxide for Multi-Modality Cancer Therapy. Biomater. Sci. 2021, 9, 3804–3813. [Google Scholar] [CrossRef] [PubMed]
  174. You, Y.; Wang, N.; He, L.; Shi, C.; Zhang, D.; Liu, Y.; Luo, L.; Chen, T. Designing Dual-Functionalized Carbon Nanotubes with High Blood–Brain-Barrier Permeability for Precise Orthotopic Glioma Therapy. Dalton Trans. 2019, 48, 1569–1573. [Google Scholar] [CrossRef] [PubMed]
  175. Zheng, M.; Liu, S.; Li, J.; Qu, D.; Zhao, H.; Guan, X.; Hu, X.; Xie, Z.; Jing, X.; Sun, Z. Integrating Oxaliplatin with Highly Luminescent Carbon Dots: An Unprecedented Theranostic Agent for Personalized Medicine. Adv. Mater. 2014, 26, 3554–3560. [Google Scholar] [CrossRef]
  176. Vinothini, K.; Rajendran, N.K.; Ramu, A.; Elumalai, N.; Rajan, M. Folate Receptor Targeted Delivery of Paclitaxel to Breast Cancer Cells via Folic Acid Conjugated Graphene Oxide Grafted Methyl Acrylate Nanocarrier. Biomed. Pharmacother. 2019, 110, 906–917. [Google Scholar] [CrossRef] [PubMed]
  177. Shao, W.; Paul, A.; Zhao, B.; Lee, C.; Rodes, L.; Prakash, S. Carbon Nanotube Lipid Drug Approach for Targeted Delivery of a Chemotherapy Drug in a Human Breast Cancer Xenograft Animal Model. Biomaterials 2013, 34, 10109–10119. [Google Scholar] [CrossRef]
  178. Liu, W.; Wei, J.; Chen, Y. Electrospun Poly(l-Lactide) Nanofibers Loaded with Paclitaxel and Water-Soluble Fullerenes for Drug Delivery and Bioimaging. New J. Chem. 2014, 38, 6223–6229. [Google Scholar] [CrossRef]
  179. Singh, S.; Mehra, N.K.; Jain, N.K. Development and Characterization of the Paclitaxel Loaded Riboflavin and Thiamine Conjugated Carbon Nanotubes for Cancer Treatment. Pharm. Res. 2016, 33, 1769–1781. [Google Scholar] [CrossRef]
  180. Burkert, S.C.; Shurin, G.V.; White, D.L.; He, X.; Kapralov, A.A.; Kagan, V.E.; Shurin, M.R.; Star, A. Targeting Myeloid Regulators by Paclitaxel-Loaded Enzymatically Degradable Nanocups. Nanoscale 2018, 10, 17990–18000. [Google Scholar] [CrossRef]
  181. Moore, T.L.; Podilakrishna, R.; Rao, A.; Alexis, F. Systemic Administration of Polymer-Coated Nano-Graphene to Deliver Drugs to Glioblastoma. Part. Part. Syst. Charact. 2014, 31, 886–894. [Google Scholar] [CrossRef]
  182. Han, M.; Zhou, Z.-H.; Luo, Y.-L.; Xu, F.; Chen, Y.-S. PH-Sensitive Carbon Nanotubes Graft Polymethylacrylic Acid Self-Assembly Nanoplatforms for Cellular Drug Release. J. Biomater. Appl. 2022, 37, 737–750. [Google Scholar] [CrossRef] [PubMed]
  183. Chowdhury, M.; Kumar Das, P. Paclitaxel-Loaded Biotinylated Fe2+ -Doped Carbon Dot: Combination Therapy in Cancer Treatment. ACS Appl. Bio Mater. 2021, 4, 5132–5144. [Google Scholar] [CrossRef] [PubMed]
  184. Xu, X.; Tang, X.; Wu, X.; Feng, X. Biosynthesis of Sorafenib Coated Graphene Nanosheets for the Treatment of Gastric Cancer in Patients in Nursing Care. J. Photochem. Photobiol. B Biol. 2019, 191, 1–5. [Google Scholar] [CrossRef] [PubMed]
  185. Ruman, U.; Buskaran, K.; Pastorin, G.; Masarudin, M.J.; Fakurazi, S.; Hussein, M.Z. Synthesis and Characterization of Chitosan-Based Nanodelivery Systems to Enhance the Anticancer Effect of Sorafenib Drug in Hepatocellular Carcinoma and Colorectal Adenocarcinoma Cells. Nanomaterials 2021, 11, 497. [Google Scholar] [CrossRef]
  186. Elsayed, M.M.A.; Mostafa, M.E.; Alaaeldin, E.; Sarhan, H.A.A.; Shaykoon, M.S.A.; Allam, S.; Ahmed, A.R.H.; Elsadek, B.E.M. Design and Characterisation of Novel Sorafenib-Loaded Carbon Nanotubes with Distinct Tumour-Suppressive Activity in Hepatocellular Carcinoma. Int. J. Nanomed. 2019, 14, 8445–8467. [Google Scholar] [CrossRef] [PubMed]
  187. Zhang, Z.; Niu, B.; Chen, J.; He, X.; Bao, X.; Zhu, J.; Yu, H.; Li, Y. The Use of Lipid-Coated Nanodiamond to Improve Bioavailability and Efficacy of Sorafenib in Resisting Metastasis of Gastric Cancer. Biomaterials 2014, 35, 4565–4572. [Google Scholar] [CrossRef]
  188. Zhang, Y.; Li, B.; Li, Z.; Xia, N.; Yu, H.; Zhang, Y. Synthesis and Characterization of Tamoxifen Citrate Modified Reduced Graphene Oxide Nano Sheets for Breast Cancer Therapy. J. Photochem. Photobiol. B Biol. 2018, 180, 68–71. [Google Scholar] [CrossRef]
  189. Chen, C.; Hou, L.; Zhang, H.; Zhu, L.; Zhang, H.; Zhang, C.; Shi, J.; Wang, L.; Jia, X.; Zhang, Z. Single-Walled Carbon Nanotubes Mediated Targeted Tamoxifen Delivery System Using Aspargine-Glycine-Arginine Peptide. J. Drug Target. 2013, 21, 809–821. [Google Scholar] [CrossRef]
  190. Yi, W.; Zhang, P.; Hou, J.; Chen, W.; Bai, L.; Yoo, S.; Khalid, A.; Hou, X. Enhanced Response of Tamoxifen toward the Cancer Cells Using a Combination of Chemotherapy and Photothermal Ablation Induced by Lentinan-Functionalized Multi-Walled Carbon Nanotubes. Int. J. Biol. Macromol. 2018, 120, 1525–1532. [Google Scholar] [CrossRef]
  191. Misra, C.; Kumar, M.; Sharma, G.; Kumar, R.; Singh, B.; Katare, O.P.; Raza, K. Glycinated Fullerenes for Tamoxifen Intracellular Delivery with Improved Anticancer Activity and Pharmacokinetics. Nanomedicine 2017, 12, 1011–1023. [Google Scholar] [CrossRef]
  192. Kumar, M.; Sharma, G.; Misra, C.; Kumar, R.; Singh, B.; Katare, O.P.; Raza, K. N-Desmethyl Tamoxifen and Quercetin-Loaded Multiwalled CNTs: A Synergistic Approach to Overcome MDR in Cancer Cells. Mater. Sci. Eng. C 2018, 89, 274–282. [Google Scholar] [CrossRef] [PubMed]
  193. Zainal-Abidin, M.H.; Hayyan, M.; Ngoh, G.C.; Wong, W.F.; Looi, C.Y. Potentiating the Anti-Cancer Profile of Tamoxifen-Loaded Graphene Using Deep Eutectic Solvents as Functionalizing Agents. Appl. Nanosci. 2020, 10, 293–304. [Google Scholar] [CrossRef]
  194. Wang, L.-H.; Sui, L.; Zhao, P.-H.; Ma, H.-D.; Liu, J.-Y.; Wei, Z.; Zhan, Z.-J.; Wang, Y.-L. A Composite of Graphene Oxide and Iron Oxide Nanoparticles for Targeted Drug Delivery of Temozolomide. Die Pharm.—Int. J. Pharm. Sci. 2020, 75, 313–317. [Google Scholar] [CrossRef]
  195. Reddy, S.; Song, L.; Zhao, Y.; Zhao, R.; Wu, D.; He, L.; Ramakrishana, S. Reduced Graphene Oxide-Based Electrochemically Stimulated Method for Temozolomide Delivery. Med. Devices Sens. 2018, 1, e10014. [Google Scholar] [CrossRef]
  196. Nandi, A.; Ghosh, C.; Bajpai, A.; Basu, S. Graphene Oxide Nanocells for Impairing Topoisomerase and DNA in Cancer Cells. J. Mater. Chem. B 2019, 7, 4191–4197. [Google Scholar] [CrossRef]
  197. Mallick, A.; Nandi, A.; Basu, S. Polyethylenimine Coated Graphene Oxide Nanoparticles for Targeting Mitochondria in Cancer Cells. ACS Appl. Bio Mater. 2019, 2, 14–19. [Google Scholar] [CrossRef] [PubMed]
  198. Wu, H.; Peng, J.; Wang, S.; Xie, B.; Lei, L.; Zhao, D.; Nie, H. Fabrication of Graphene Oxide-β-Cyclodextrin Nanoparticle Releasing Doxorubicin and Topotecan for Combination Chemotherapy. Mater. Technol. 2015, 30, 242–249. [Google Scholar] [CrossRef]
  199. Au, T.H.; Nguyen, B.N.; Nguyen, P.H.; Pethe, S.; Vo-Thanh, G.; Vu Thi, T.H. Vinblastine Loaded on Graphene Quantum Dots and Its Anticancer Applications. J. Microencapsul. 2022, 39, 239–251. [Google Scholar] [CrossRef]
  200. Li, N.; Zhao, Q.; Shu, C.; Ma, X.; Li, R.; Shen, H.; Zhong, W. Targeted Killing of Cancer Cells in Vivo and in Vitro with IGF-IR Antibody-Directed Carbon Nanohorns Based Drug Delivery. Int. J. Pharm. 2015, 478, 644–654. [Google Scholar] [CrossRef]
  201. Georgakilas, V.; Otyepka, M.; Bourlinos, A.B.; Chandra, V.; Kim, N.; Kemp, K.C.; Hobza, P.; Zboril, R.; Kim, K.S. Functionalization of Graphene: Covalent and Non-Covalent Approaches, Derivatives and Applications. Chem. Rev. 2012, 112, 6156–6214. [Google Scholar] [CrossRef]
  202. Calabrese, G.; De Luca, G.; Nocito, G.; Rizzo, M.G.; Lombardo, S.P.; Chisari, G.; Forte, S.; Sciuto, E.L.; Conoci, S. Carbon Dots: An Innovative Tool for Drug Delivery in Brain Tumors. Int. J. Mol. Sci. 2021, 22, 11783. [Google Scholar] [CrossRef] [PubMed]
  203. Nocito, G.; Calabrese, G.; Forte, S.; Petralia, S.; Puglisi, C.; Campolo, M.; Esposito, E.; Conoci, S. Carbon Dots as Promising Tools for Cancer Diagnosis and Therapy. Cancers 2021, 13, 1991. [Google Scholar] [CrossRef] [PubMed]
  204. Das, P.; Ganguly, S.; Saravanan, A.; Margel, S.; Gedanken, A.; Srinivasan, S.; Rajabzadeh, A.R. Naturally Derived Carbon Dots In Situ Confined Self-Healing and Breathable Hydrogel Monolith for Anomalous Diffusion-Driven Phytomedicine Release. ACS Appl. Bio Mater. 2022, 5, 5617–5633. [Google Scholar] [CrossRef]
  205. Consoli, G.M.L.; Giuffrida, M.L.; Zimbone, S.; Ferreri, L.; Maugeri, L.; Palmieri, M.; Satriano, C.; Forte, G.; Petralia, S. Green Light-Triggerable Chemo-Photothermal Activity of Cytarabine-Loaded Polymer Carbon Dots: Mechanism and Preliminary In Vitro Evaluation. ACS Appl. Mater. Interfaces 2023, 15, 5732–5743. [Google Scholar] [CrossRef]
  206. Forte, G.; Consiglio, G.; Satriano, C.; Maugeri, L.; Petralia, S. A Nanosized Photothermal Responsive Core-Shell Carbonized Polymer Dots Based on Poly(N-Isopropylacrylamide) for Light-Triggered Drug Release. Colloids Surf. B Biointerfaces 2022, 217, 112628. [Google Scholar] [CrossRef]
  207. Liu, Y.; Zhao, Y.; Sun, B.; Chen, C. Understanding the Toxicity of Carbon Nanotubes. Acc. Chem. Res. 2013, 46, 702–713. [Google Scholar] [CrossRef] [PubMed]
  208. Saleemi, M.A.; Hosseini Fouladi, M.; Yong, P.V.C.; Chinna, K.; Palanisamy, N.K.; Wong, E.H. Toxicity of Carbon Nanotubes: Molecular Mechanisms, Signaling Cascades, and Remedies in Biomedical Applications. Chem. Res. Toxicol. 2021, 34, 24–46. [Google Scholar] [CrossRef]
  209. Kolosnjaj, J.; Szwarc, H.; Moussa, F. Toxicity Studies of Fullerenes and Derivatives. In Bio-Applications of Nanoparticles; Advances in Experimental Medicine and Biology; Chan, W.C.W., Ed.; Springer: New York, NY, USA, 2007; pp. 168–180. ISBN 978-0-387-76713-0. [Google Scholar]
Figure 1. Number of anticancer drugs according to class.
Figure 1. Number of anticancer drugs according to class.
Pharmaceutics 15 01545 g001
Figure 2. A stacked bar chart depicting the number of database entries for each CNM type under each class of anticancer drug.
Figure 2. A stacked bar chart depicting the number of database entries for each CNM type under each class of anticancer drug.
Pharmaceutics 15 01545 g002
Figure 3. The number of entries that have DLEs, DLCs, or DREs above or below 50%, grouped by CNM class.
Figure 3. The number of entries that have DLEs, DLCs, or DREs above or below 50%, grouped by CNM class.
Pharmaceutics 15 01545 g003
Table 1. Database of CNM-based anticancer nanocarriers.
Table 1. Database of CNM-based anticancer nanocarriers.
Chemotherapeutic—Drug ClassCNM-Based
Nanocarrier
Biological Study Models Drug Loading and
Release Metrics
Experimental ResultsRef.
Pharmaceutics 15 01545 i001
5-fluorouracil (5-FU)—
pyrimidine antimetabolite
CS-carbon quantum dot (CQD)-Aptin vitro: MCF-7 cells32% DLC, approximately 100% DRE, and pH-sensitive, controlled 5-FU releaseThe unloaded nanocarrier is biocompatible, and the use of an aptamer increases uptake and cytotoxicity in breast cancer cells.[26]
FA-PEG-bis-amine multiwalled carbon nanotube (MWCNT)in vitro: MCF-7 cells99% DLE, ~90% DRE, with pH-triggered drug release sustained over 900 minThis nanocarrier increases circulation time, half-life, and accumulation of 5-FU in target tissues, and this leads to the effective killing of breast cancer cells in vitro.[27]
CS/Au/MWCNTin vitro: MCF-7 cells43% DLC, 59% DRE, with prolonged, sustained drug releaseReduced potential side effects and increased efficacy compared to free 5-FU were observed. A reduction in cancer cell viability was observed at low nanocarrier concentrations.[28]
Nanodiamond (ND)-ADHin vitro: MCF-7 and HepG2 cells88% DLE, 35% DRE, with pH-mediated, sustained drug releaseThis nanocarrier showed potent anticancer effects with low haemolytic toxicity in human blood.[29]
Mesoporous carbon nanoframe (mCNF)in vitro: HeLa cells31% DLC, 80% DRE, with dual pH/NIR-triggered drug releaseThis system displayed excellent photothermal efficiency with the NIR pulse-triggered burst release of 5-FU. The photothermal conversion efficiency of this system was found to be 21%. This synergistic chemo–photothermal therapy combined with photoacoustic imaging capabilities can effectively treat cancer in vitro.[30]
PEG-C60
fullerene–alanine
in vitro: MCF-7 and BGC-823 cells1% DLC, with no quantitative drug release dataThe unloaded nanocarrier displays good biocompatibility and the system is stable in murine serum for over 24 h. This formulation results in the significantly better inhibition of cancer cells compared to free 5-FU.[31]
Graphene oxide (GO)in vitro: A549 cells31% DLC, 35% DRE, with pH-triggered drug releaseThe blank nanocarrier is biocompatible and the loaded system improved the stability of 5-FU.[32]
CS-coated Fe3O4–NH2/graphene quantum dot (GQD) nanohybridin vitro: A549 cells90% DLC, 84% DRE, with pH-dependent drug releaseThis system has magnetic resonance/fluorescence imaging capabilities and displayed significantly higher cytotoxicity than free 5-FU, whilst the unloaded nanocarrier is biocompatible.[33]
HPMC/GOin vitro: Vero, HepG2, and A549 cells No quantitative drug loading/release studies were performed. The blank nanocarrier displays high biocompatibility in normal cells, whilst the drug-loaded system displays a higher antitumour efficacy than free 5-FU. A green synthesis method was used.[34]
TAU-GOin vitro: HepG2 cells; in vivo: SD rats50% DLE, 90% DRE, with pH-triggered 5-FU releaseThis biocompatible nanocarrier improved the circulation time and anticancer efficacy of 5-FU.[35]
Carbon dot (CD)-BTin vitro: MCF-7, HeLa, and HEK-296 cells35% DLE, 81% DRE, with pH-triggered drug releaseAn initial burst of 5-FU is followed by sustained release; this nanocarrier also displays fluorescence imaging capabilities. BT-mediated targeting of cancer cells resulted in high cytotoxicity towards neoplastic cells and increased cellular uptake due to biotin-receptor-mediated endocytosis.[36]
N-doped mesoporous carbon sphere (NMCS)-DSPE-PEGin vitro: B16F0 cells38% DLC, 78% DRE, with dual pH/NIR-triggered drug releaseThis nanocarrier produces reactive oxygen species when irradiated with an NIR laser, and the resulting PDT/PTT/chemotherapeutic combination therapy effectively kills melanoma cells much more efficiently than 5-FU alone.[37]
Pharmaceutics 15 01545 i002
6-mercaptopurine (6-MP)—
purine antimetabolite
CD-BTin vitro: CHO, MCF-7, and HepG2 cells5% DLC, 79% DRE, with dual pH- and redox-sensitive drug releaseComparable anticancer activity to free 6-MP (in cancer cells) with much lower cytotoxicity (in healthy cells). A GSH-sensitive carbonyl vinyl sulphide group was used to bind 6-MP to BT.[38]
Pharmaceutics 15 01545 i003
Anastrozole—aromatase
inhibitor
GO-Fe nanoparticlesin vitro: MCF-7 cells84% DLENo qualitative drug release data shown. The system displays higher cytotoxicity than the free drug for cancer cells, and it has magnetic properties.[39]
Pharmaceutics 15 01545 i004
Bortezomib—proteasome
inhibitor
CD-CuS NPs-MMT7in vitro: U251 MG cells Synergistic drug delivery and PTT platform that specifically targets cancer cells. No qualitative drug loading/release data shown. This innovative nanocarrier combines immune system evasion capabilities with the enhanced suppression of tumour growth and metastasis to provide excellent control over cancer growth and metastasis.[12]
Pharmaceutics 15 01545 i005
Capecitabine—pyrimidine
antimetabolite
Single-walled carbon nanotube (SWCNT)-FL-FA-NCCin vitro: Caco-2/TC7 cells No quantitative drug loading/release data shown. This nanocarrier is nontoxic and has fluorescence imaging capabilities. The effective targeting of colon cancer cells leads to an increase in anticancer activity compared to the free drug.[40]
oxiSWCNT-CS-FAin vitro: COLO320DM and HT29 cells; in vivo: albino rabbits94% DLE, 89% DREAn increase in cytotoxicity compared to free drug was noticed during in vitro experiments. The capsule formulation of this nanocarrier is exclusively released in the colon in vivo, avoiding premature release in the stomach. Active targeting of cancer cells was achieved via the FA-targeting ligand.[41]
Pharmaceutics 15 01545 i006
Carboplatin (CP)—DNA
alkylating agent
GO-PAMAMin vitro: hMSC and HeLa cells No quantitative drug loading/release data shown.
The 100 nm width GO (unloaded) was found to be the least toxic. This system displayed enhanced anticancer activity compared to free CP, with decreased cytotoxicity.
[42]
GO-gelatinein vitro: IMR-32 and hMSC cells99% DLE, with no quantitative drug release data shownThis formulation displays effective CP delivery and uptake in vitro, resulting in a higher potency than free CP. Excellent biocompatibility and stability were observed in vitro.[43]
oxiMWCNT-HAin vitro: TC–1 and NIH/3T3 cellsNo quantitative drug loading/release data shownThis system displayed the selective uptake and targeting of cancer cells over healthy cells, resulting in significantly higher cytotoxic effects in neoplastic cells and lower side effects in healthy cells.[44]
Aminated MWCNTin vitro: MDA-MB-23 and MCF-12A89% DLC, 21% DREThis formulation provided increased cancer cell death compared to free CP and killed cells via an ROS-triggered autophagy mechanism.[45]
FA-CDT-C60 fullerenein vitro: HeLa, HeLa-RFP, and A549 cells; in vivo: Danio rerio, both healthy and bearing HeLa tumours37% DLC, ~80% DRE, with pH-triggered drug releaseThis system displayed increased anticancer effects compared to the free drug alone due to the active targeting of folate-receptor-overexpressing cancer cells and improved cellular uptake. Low toxicity and improved antitumour effects compared to the free drug were also seen in vivo.[46]
CS-Fe3O4-GOin vitro: HepG2 and MCF-7 cells74% DLE, 90% DRE, with pH-triggered drug releaseA very high amount of CP was released at neutral pH.
Despite this, an increase in CP potency and a reduction in systemic toxicity was observed.
[47]
GO-CS-FAin vitro: LX-2 and SKOV3 cells14% DLC, ~90% DRECP release was similar in neutral and acidic environments; hence, this system is unsuitable for pH-triggered drug release via noncovalent drug attachment. The system showed slightly lower cancer cell inhibition than free CP.[48]
GO-Fe3O4-PANIin vitro: SMMC-7721, HepG2, and HL-7702 cells~95% DRE; the qualitative drug loading data provided does not account for unbound CP that was removedThe blank nanocarrier showed efficient cellular uptake and negligible cytotoxicity. This nanocarrier has magnetic properties and pH-triggered drug release.[49]
Pharmaceutics 15 01545 i007
Carmustine—DNA
alkylating agent
GO-PAAin vitro: GL261 cells19% DLC, with no quantitative drug release data shownA significant increase in half-life, >70% decrease in IC50 value, and 30% increase in inter-strand DNA crosslinking was observed compared to the free drug in vitro.[50]
N-doped carbon nanotube (CNT) spongesin vitro: rat astrocytes, C6, RG2, and U87 cells ~90% DRE, with no quantitative drug loading dataThis nanocarrier displayed similar cytotoxicity to the free drug, with a sustained-release profile. The sponges appear to be more biocompatible than CNTs alone; hence, the blank nanocarrier showed low cytotoxicity, whilst the drug-loaded system displayed strong anticancer effects.[51]
Pharmaceutics 15 01545 i008
Chlorambucil—DNA
alkylating agent
Reduced graphene oxide (rGO)-FA-gelatinein vitro: Siha cells35% DLC, 82% DRE, with pH-triggered drug releaseA significant decrease in IC50 value compared to the free drug was observed. The use of gelatine facilitated sustained drug release. This system is a promising treatment for cervical adenocarcinoma.[52]
Pharmaceutics 15 01545 i009
Cisplatin (CisP)—DNA
alkylating agent
GO-Alain vitro: MCF-7 and HepG2 cells4% DLC and ~70% DRE, with sustained drug release at neutral pHThe blank nanocarrier is biocompatible, whilst the CisP-loaded material is effective at killing cancer cells in vitro.[53]
Oxidised carbon nanohorn (CNH)in vitro: NCI-H460Approximately 1% DLC and 80% DRE, with sustained CisP releaseThis nanosystem displayed similar anticancer effects to the free drug in vitro.[54]
CNH-quantum dotin vitro: AY-27 cells18% DLC and 70% DRE, with a sustained drug release profileThis theragnostic system displayed a significant reduction in anticancer potency compared to free CisP; however, it has imaging capabilities arising from the inclusion of CdSe quantum dots.[55]
Ultra-short SWCNTin vivo: SCID/beige mice bearing MCF-7, BCM-4272, and MDA-MB-231 tumours This nanoformulation effectively treated CisP-resistant breast cancer in a xenograft mouse model. The nanocarrier also displays an enhanced circulation time and increased tumour localisation, leading to increased potency compared to the free drug.[56]
Silane-modified NDin vitro: HeLa cells This unique system has a Pt loading of 0.25 mmol/g ND, and CisP is not released from the conjugate. Despite this, the system displayed a similar IC50 value to CisP. The main advantage of this system is the prevention of CisP isomerisation, leading to enhanced aqueous stability. No quantitative drug loading/release data were given.[15]
SA/NDin vitro: HepG2, HeLa, A549, and RAW264.7 cells This sustained-release drug platform improved the CisP accumulation in cancer cells, with improved drug safety. Whilst no quantitative drug loading or release data were given, no change in the antitumour mechanism was observed compared to the free drug.[57]
EGF-NDin vitro: HepG2 cells1% DLC, with no quantitative drug release data shownThis nanoformulation was capable of selectively killing liver cancer cells and displayed increased potency compared to free CisP. This was due to the EGF-mediated targeting of cancer cells. In addition, the NDs are probes for 3D Raman microscopy imaging. This nanocarrier system induces morphological changes in cancer cells, resulting in higher surface areas for CisP absorption with a lower risk of adverse side effects.[58]
C60 fullerenein vivo: BALB/c mice50% DLC, with no quantitative drug release data shownA two-fold decrease in systemic toxicity (LD50) compared to free CisP was observed. Specifically, the nanocarrier decreased drug-induced leukopenia, anaemia, thrombocytosis, and inflammation.[59]
C60 fullerenein vitro: LLC cellsNo quantitative drug loading or release data givenA 4.5× decrease in IC50 value compared to the free drug was observed in vivo. The fullerene itself was found to increase the cellular uptake and accumulation of CisP.[60]
oxiC60 fullerenein vitro: L929 cells16% DLC, 60% DRE, with pH-triggered drug releaseThis nontoxic nanocarrier displayed outstanding fluorescence properties for cellular imaging experiments.[61]
C60 fullerenein vitro: HCT-116, HeLa, HL-60, HL-60/adr, and HL-60/vinc cells; in vivo: C57BL/6J mice bearing LLC tumoursNo quantitative drug loading/release data shownThis nanocarrier killed chemotherapy-resistant leukaemia cells in vitro and exhibited effective lung cancer tumour growth inhibition in vivo. Molecular docking studies suggested that the fullerene binds to proteins involved in chemotherapy resistance.[62]
CS-GOin vitro: HeLa cells71% DLE, 88% DRE, with pH-triggered drug releaseThe functionalisation of GO with CS and CisP dramatically reduced protein binding. This biocompatible nanocarrier triggered apoptosis in drug-resistant cancer cells.[63]
CQD-GE11-DOXin vitro: CNE-2 cells; in vivo: BALB/c mice bearing CNE-2 tumours5% DLC and 57% DRE, with pH-triggered drug releaseSpecific tumour targeting and inhibition was observed in vivo, and the effective killing of nasopharyngeal carcinoma cells was exhibited in vitro. This nanocarrier has fluorescence imaging capabilities and showed no obvious side effects in vivo.[64]
GO-PEGin vitro: MG63, SAOS-2, U2-OS, MDA-MB-231, MDA-MB468, U118, and U87 cells64% DLE, with no quantitative drug release data shown and redox-sensitive drug delivery using a CisP prodrugThis nanoformulation displayed high uptake and proliferation inhibition in osteosarcoma cells, and effective internalisation, but reduced potency in glioblastoma cells. This system is capable of inhibiting cell migration in highly invasive breast carcinoma.[65]
MnO2-GO-Ce6-HAin vitro: MDA-MB-231 and RLE-6TN cells; in vivo: BALB/c mice bearing MDA-MB-231 tumours7% DLC and 60% DRE, with pH-triggered drug releaseCombination therapy of (1) MnO2 to regulate the tumour microenvironment, enhancing the anticancer effect of (2) CisP chemotherapy and (3) Ce6 PDT. The incorporation of HA facilitates tumour targeting for a true theragnostic system. This system also shows excellent biocompatibility and antitumour efficacy in vivo.[66]
rGO-PHEMA-DOXin vitro: MCF-7 cells82% DLE and 64% DRE, with pH-triggered drug releaseA significant decrease in IC50 value compared to free CisP and DOX was observed. This biocompatible nanocarrier displayed efficient cellular uptake and a synergistic effect between the two loaded drugs, resulting in the effective killing of breast cancer cells.[67]
Fe3O4-rGO-PHEMA-METin vitro: HepG2 and Caco-2 cells; in vivo: BALB/c mice, both healthy and bearing HepG2 tumours82% DLE, 60% DRE, with pH-triggered drug releaseNo side effects and potent antitumour efficacy was noted in vivo. This highly biocompatible nanosystem effectively killed hepatocellular carcinoma in vitro.[68]
GO-PEG-DOXin vitro: CAL-27, L929, and MCF-7 cells; in vivo: nude mice carrying CAL-27 tumours37% DLC and 65% DRE, with pH-triggered drug releaseA 2× increase in cancer cell apoptosis and necrosis compared to the single-drug-loaded nanocarrier was observed. An attenuation of toxicity and enhanced anticancer effects compared to free DOX/CisP were observed.[69]
S-doped CDin vitro: A2780 and A2780 cellsNo quantitative drug loading/release data shownThe unloaded CDs were found to be biocompatible and could interact with proteins and lipids on the surfaces of cancer cells. A similar IC50 value to the free drug was seen in normal ovarian cancer cells, and the nanoformulation could kill drug-resistant cancer cells.[70]
CD-iRGDin vitro: A549, HUVEC, and HEL-299No quantitative drug loading/release data shownThis nanocarrier destroyed lung cancer cells whilst leaving healthy cells unharmed.[71]
MWCNTin vitro: A549 and A549/DDP cells; in vivo: BALB/c mice carrying A549/DDP tumoursNo quantitative drug loading/release data givenThe unloaded MWCNTs were found to be biocompatible, whilst the loaded nanocarrier had higher cytotoxicity against cancer cells than free CisP. This nanoformulation could effectively treat a drug-resistant lung cancer in vivo model.[72]
PDA CD-anti-EpPCANin vitro: HepG2 cells; in vivo: BALB/c mice bearing HepG2 tumours This synergistic nanocarrier combined a cisplatin prodrug with significant PTT and fluorescence imaging capabilities for effective image-guided chemo–photothermal therapy. This biocompatible system exhibited excellent antitumour effects in vitro and in vivo.[73]
CD-PEGin vitro: GES-1 and MGC-803 cells5% DLC, with no quantitative drug release data given and pH/redox-mediated drug release achieved via a hydrolysable benzoic imine bondA CisP prodrug was used, and the resulting system had comparable anticancer efficacy to free CisP, with reduced side effects. This system also exhibited fluorescence imaging capabilities.[74]
CD-PEGin vitro: A549, HUVEC, and HEL-299 cellsNo quantitative drug loading data given, 86% DRE and redox-sensitive drug releaseThis fluorescent nanocarrier could effectively kill cancer cells whilst leaving healthy cells unharmed.[75]
Pharmaceutics 15 01545 i010
Curcumin (CUR)—ROS scavenger and lipid peroxidation inhibitor
Fluorinated GO-LINin vitro: MCF-7 and MCF10A cells; in vivo: BALB/c mice with 4T1-induced breast cancer61% DLC, 95% DRE, with pH-sensitive drug releaseThis simple, low-cost system acts as a contrast agent for magnetic resonance imaging. It also displays improved cytotoxicity, tumour suppression, and reduced side effects compared to free CUR due to effective cancer cell targeting.[76]
GQD-GlcNin vitro: MCF-7 cellsNo quantitative drug loading information, 37% DRE with pH-triggered, sustained drug releaseThe unloaded nanocarrier is biocompatible, whilst the loaded system exhibits effective cancer cell targeting and internalisation in vitro. This system also has fluorescence imaging capabilities.[77]
SWCNTin vitro: PC-3 cells94% DLE, 95% DRE, with pH-triggered drug releaseThe combination of efficient CUR delivery and SWCNT-mediated PTT successfully inhibited tumour cell growth. This nanocarrier also reduced CUR biodegradation and increased its solubility.[78]
SWCNT-PC-PVPin vitro: PC-3 and S180 cells; in vivo: Kunming mice bearing S180 tumours No quantitative drug loading/release data. This biocompatible nanocarrier increased CUR cellular uptake, plasma concentration, and bioavailability. The system overcomes the main barrier to the low anticancer effect of free CUR (low plasma concentration) whilst displaying low in vivo toxicity. This is a combination therapy with the SWCNT-mediated photothermal ablation of cancer cells.[79]
oxiND-ADHin vitro: MCF-7 and
HepG2 cells
93% DLE, 36% DRE, with pH-triggered, sustained drug releaseThe use of a pH-sensitive amide bond to bind CUR slows release and increases stability, resulting in potent cytotoxicity.[29]
Graphene oxide quantum dot (GOQD)-CS-PEG-MUC-1 aptamerin vitro: MCF-7 and HT-2 cells99% DLC, 64% DRE, with pH-responsive drug releaseThis system effectively targets MUC-1-overexpressing cancer cells whilst displaying photoluminescence imaging and cancer detection abilities. An increase in therapeutic efficacy and cellular uptake compared to free CUR and low haemolysis with human blood was observed with this system.[19]
CD-PNMin vitro: SH-SY5Y cellsNo qualitative CUR loading information, 82% DREThis formulation resulted in a 10× enhancement of CUR solubility whilst displaying excellent photophysical properties and low toxicity.[80]
CDin vitro: HepG2 and A549 cells3% DLC, ~90% DRE, with pH-mediated drug releaseThis cost-effective, photoluminescent nanocarrier is nontoxic to normal cells and displayed potent anticancer effects with enhanced CUR bioavailability and a small size.[81]
CoFe2O4/GO-ADH-CMCin vitro: MDA-MB-231 and MCF-10A cells2% DLC, 86% DRE, with pH-triggered, controlled drug releaseA decrease in cancer cell viability compared to free CUR was noted when using this system.[82]
GQD-HAin vitro: HeLa and L929 cells98% DLE, ~100% DRE, with pH-triggered drug releaseThis nanoformulation displayed excellent anticancer activity compared to CUR alone and has no toxic effect on healthy cells. The system is also highly fluorescent when CUR is released.[83]
GO-BSA-AS1411 aptamerin vitro: MCF-7 and SKBR3 cells9% DLC, 70% DRE, with pH-triggered drug releaseEfficient targeting of nucleolin-overexpressing MCF7 cancer cells was achieved, facilitated by aptamer attachment. This resulted in increased CUR antitumour activity. BSA decoration was found to improve nanocarrier biostability and slow CUR degradation.[84]
CS-Fe3O4-RGOin vitro: MCF-7 cells95% DLE, 96% DREThis system successfully targeted and induced apoptosis in breast cancer cells. The superparamagnetic nanocarrier increased the rate of CUR delivery compared to the free drug.[85]
Pharmaceutics 15 01545 i011
Cyclophosphamide—DNA alkylating agent
Hollow mesoporous carbon sphere (HMCS)in vitro: CNE cells; in vivo: nude mice with CNE tumours20% DLC, with no release experiments carried out with PTXThe HMCSs themselves exhibited strong antitumour effects through PTT.[86]
PAA/PEG/CNT/MTX Approximately 80% DRE, with no quantitative drug loading data shownThe system displayed dual pH- and temperature-triggered drug release, with an initial burst of drug followed by sustained delivery.[87]
Pharmaceutics 15 01545 i012
Cytarabine—DNA polymerase inhibitor
GQD-CS 73% DLE, 72% DRE, with pH-sensitive drug release via a hydrolysable amide linkageThis nanosystem possesses remarkable fluorescence stability, and CS wrapping enhanced the water solubility of this system.[88]
Au/GQD/MPA/PEIin vitro: HL-60 cells68% DLE, 78% DRE, with dual pH- and NIR-triggered drug releaseCytarabine was attached via charge–dipole interactions. The chemo–photothermal combination therapy had higher efficacy than PTT alone.[21]
Pharmaceutics 15 01545 i013
Dabrafenib—reversible ATP-competitive kinase inhibitor
GO-BSAin vitro: A375, HDF, SKmel28, SKmel23, MelJuSo, MNT-1, and NHEM cellsNo quantitative drug loading/release data, with pH-triggered drug releaseThe potency of dabrafenib was retained, with effective BRAF and HDAC inhibition in human melanoma cells.[89]
Pharmaceutics 15 01545 i014
Dasatinib—tyrosine kinase
inhibitor
PLA-PGA-PEG-CNTin vitro: U-87 cells4% DLC, ~65% DREThis nanocarrier system was synthesised via a simple one-pot method and demonstrated improved therapeutic efficacy compared to the free drug in vitro. The drug release profile of this system can be controlled by varying the composition of the polymer coating.[90]
Pharmaceutics 15 01545 i015
Daunorubicin (DNR)—
topoisomerase II inhibitor
P-gp-GGNin vitro: adriamycin-resistant leukaemia cell lines KA and K562/A02 cells; in vivo: KA nude mice with drug-resistant leukaemia-cell-induced tumours32% DLC, roughly 45% DRE, with redox-triggered DNR release facilitated by increasing glutathione concentrationThis nanocarrier overrides the cell‘s drug resistance to facilitate DNR uptake, resulting in a remarkable inhibition of tumour growth in vivo.[20]
PLA/MWCNT/FE3O4in vitro: K562 cells96% DLE, roughly 55% DRE, with dual magnetic field- and pH-mediated drug releaseThe most effective killing of leukaemia cells was observed at a 20 µg/mL nanocarrier concentration.[91]
NDin vitro: K562 cells~95% DLC, most of the bound DNR was releasedThis formulation achieved a three-fold reduction in the IC-50 value compared to free DNR for the treatment of drug-resistant K562 cells.[92]
f-CNTs 94% DLE at 3:1 DNR:f-CNT ratio, with no quantitative drug release data providedHydroxylated CNTs provided the best DNR binding (through electrostatic interactions).[93]
Bi2MoO6/NH2-GO/PEGin vitro: HUVEC and MCF-7 cells33% DLC, 86% DRE, with pH-triggered drug releaseDNR was selectively released in cancer cells. Haemolysis and coagulation tests prove system has no negative effects on the blood.[94]
Pharmaceutics 15 01545 i016
Decitabine—DNA methyltransferase inhibitor
A1-GOin vitro: A549, NCI-H157, NCI-H520, NCI-H1299, NCI-H446, MCF-7, and HeLa cells64% DLE, 75% DRE, with pH-dependent drug releaseSpecific recognition and targeting of lung cancer cells over other cancer cells was achieved by using the A1 aptamer. This system achieved much higher anticancer efficacy than the free drug.[95]
Pharmaceutics 15 01545 i017
Docetaxel (DTX)—microtubule growth inhibitor
GO-PEGin vitro: DU-145 cellsNo quantitative drug loading or release data providedThis system was highly effective at killing prostate cancer cells due to a decrease in IC50 compared to free DTX. The nanocarrier displayed low dispersion stability in biological fluids.[96]
RGD-CS-SWCNTin vitro: A549 and MCF-7 cells; in vivo: BALB/c mice inoculated with A549 tumours32% DLC, 68% DRE, with pH-triggered drug releaseSignificant drug uptake and growth inhibition in A549 cells was observed. The system entered cells via clathrin- and caveolin-mediated endocytosis and displayed strong tumour targeting, growth inhibition, and biosafety in vivo.[97]
Oxi-carbon nano-horn (CNH)-PEG-mAbin vitro: MCF-7; in vivo: ICR mice xenografted with H22 tumours74% DLE, 59% DREThe adsorption of DTX to the nanohorns was achieved via π–π stacking. Prolonged diffusion-controlled DTX release was achieved. The use of mAb resulted in the selective killing of cancer cells in vitro and in vivo and a lower IC50 and no significant side effects compared to free DTX in vivo. This nanocarrier also leveraged the enhanced permeability and retention effect.[98]
Acylated C60 fullerenein vitro: MCF-7 and MDA-MB-23 cells; in vivo: Wistar rats81% DLC, 84% DREThis system achieved 4.2× higher bioavailability and 50% lower drug clearance compared to free DTX. This resulted in enhanced cancer cell cytotoxicity, low haemolysis, and high erythrocyte compatibility.[99]
Hexagonal nanostructured GOin vitro: A549 cells41% DLE and approximately 20% DRE, with pH-mediated drug releaseThe nanostructured material improved the drug loading capacity compared to pristine GO and displayed good biocompatibility.[32]
Carbon nanoparticle (CNP)-HIF-PLGAin vitro: Walker256 cells; in vivo: rats xenografted with Walker256 tumours16% DLC, with no qualitative release data shown and NIR-activated drug releaseThis nanocarrier displayed photothermal properties. The synergistic effect of chemotherapy and image-guided NIR PTT gave this system the ability to effectively target and treat metastatic lymph nodes both in vitro and in vivo, with minimal side effects.[100]
CNP-PLGAin vitro: MDA-MB-231 and HUVEC cells; in vivo: New Zealand white rabbits bearing VX2 liver tumoursNIR-triggered drug release, with no quantitative drug loading/release data shownThis system relied on a combination of PTT and photoacoustic imaging to treat cancer in vitro and in vivo. Highly targeted drug delivery was achieved by transport through the lymphatic system to produce an excellent therapeutic effect on metastatic lymph nodes with favourable biocompatibility and biosafety.[101]
CNSin vitro: MDA-MB 231 cells92% DLE, with no quantitative release data shownCS nanopores substantially assisted in drug loading to give this system favourable anticancer properties.[102]
Hydroxylated CNT-APAin vitro: MDA MB-231 cells; in vivo: Wistar rats51% DLE, with no quantitative drug release data shown and pH-triggered drug release in cancer cytosolThis formulation achieved a 2.8× enhancement in cytotoxicity and superior pharmacokinetics compared to free DTX, with substantial hemocompatibility with human blood and reduced side effects compared to the drug alone in vivo.[103]
RBC@GQDin vitro: A549 cells; in vivo: CAnN.CgFoxn mice carrying A549-induced tumours; C57BL/6J mice with ALTS1C1 intracranial tumoursApproximately 40% DRE, with no qualitative drug loading data and NIR-triggered drug releaseThis system achieved an eight-fold increase in the accumulation in tumour tissues compared to the free drug. The synergy between the chemotherapy and photolytic properties of the nanocarrier allowed for deep penetration into tumours and effective treatment in vivo.[104]
C60 fullerene-APAin vitro: MDA MB-231 cells; in vivo: Wistar rats48% DLE, 96% DRE, with pH-triggered drug releaseA substantial decrease in haemolysis (human blood) and protein binding (BSA) compared to free DTX was observed. The nanoformulation also increased bioavailability and potency compared to free DTX. Fullerenes display partial P-gp efflux inhibition.[105]
Pharmaceutics 15 01545 i018
Doxorubicin (DOX)—
topoisomerase II inhibitor
GO/PEI.Ac-FITC-PEG-LAin vitro: SMMC-7721 and PIEC cells85% DLC, ~80% DRE, with pH-dependent drug releaseThe nanocarrier shows specificity for ASGPR-receptor-containing cancer cells whilst retaining DOX therapeutic efficacy.[106]
GO-PRM/SAin vitro: MCF-7 cells 29% DLC, 49% DRE, with pH-dependent drug releaseProtein adsorption in physiological environments was suppressed and the system showed enhanced cytotoxicity compared to GO-DOX alone.[107]
Tf/FA/GO/PF68in vitro: SMMC-7721 and L-02 cells96% DLC, 55% DRE, with pH-dependent drug releaseThe nanocarrier displayed low toxicity and high specificity due to the synergistic effect of Tf- and FA-targeting ligands on cancer cell targeting. The DOX-loaded nanosystem was not tested on healthy L-02 cells.[108]
GNRs/GO@PDA nanosheetsin vitro: MCF-7 cells81% DLE, 49% DREThis system displayed dual pH/NIR-responsive drug release. The GNRs act as a probe for PTT.[109]
CNH/DCA-HPCHSin vitro: 4T1 cells; in vivo: mice bearing 4T1 tumours This system combines chemotherapy with NIR-mediated PTT for the treatment of cancer in vitro and in vivo. The nanocarrier displays pH-dependent drug release; however, no DOX quantification for drug loading/release was included.[110]
oxiCNH/SA-mAbin vitro: MCF-7 and HEK293 cells; in vivo: CR male mice bearing subcutaneous hepatic H22 tumours100% DLC, 67% DRE, with pH-dependent drug releaseDOX is released in the endosomes of MCF-7 cells. Specific targeting of VEGF-containing cancer cells over healthy HEK293 cells was achieved by using mAb as a targeting ligand. This nanocarrier was more effective than free DOX both in vivo and in vitro, whilst showing reduced liver and cardiac toxicity.[9]
PEG-SWCNTin vivo: SCID mice bearing Raji lymphoma xenografts100% DRE, with pH-dependent drug releaseIncreased tumour inhibition and reduced systemic toxicity compared to free DOX was determined through in vivo experiments.[111]
C60 fullerenein vitro: CCRF-CEM, Jurkat, THP1, and Molt-16 cells Long-term nanocarrier stability was observed in physiological saline solution. Complexation with C60 fullerene promoted Dox entry into leukemic cells, resulting in ≤ 3.5 higher cytotoxicity compared to free DOX.[112]
Fullerenolin vitro: MCF-7 and
MDA-MB-231 cells; in vivo: zebrafish embryo (Danio rerio, wild type)
This formulation resulted in enhanced uptake, decreased proliferation, and remarkable cytotoxicity of DOX in breast cancer cells. A decrease in zebrafish embryotoxicity compared to DOX alone was noted. No drug loading or release analysis was included in this study.[113]
C82 fullerene-cRGDin vitro: NCl-H2135 cellsRedox-dependent DOX release, triggered by increasing glutathione concentrationSignificant cytotoxicity at low doses was observed due to enhanced cellular uptake relative to free DOX.[114]
DSPE-PEG-NDin vitro: 4T1 cells; in vivo: BALB/c mice injected with 4T1 cells and SD ratsApproximately 95% DLE and 34% DREFavourable circulation time, accumulation in tumour tissues, and ability to deliver DOX to tumour cell nucleus was observed. This leads to a significant enhancement of DOX efficacy and biocompatibility.[115]
ND-PEGin vitro: A549 cells65% DLC, approximately 60% DREThe nanocarrier was readily soluble in PBS and water, and unloaded ND-PEG displayed excellent biocompatibility. This system was able to deliver DOX directly to cancer cells.[116]
Carbon nanoring (CNR)in vitro: A549, Hela, L929, and BEAS-2B cells51% DLE, approximately 80% DRE, with pH-mediated drug releaseEnhanced antitumor activity compared to free DOX was achieved. The DOX-CNR system was highly selective, with much higher cytotoxicity for cancerous cells than normal cells. DOX was released in the nuclei of cancer cells.[117]
GA-MWCNT
MA-MWCNT
in vitro: MDA-MB-231 and MCF-7 cells97% DLE and 71% DRE for MA-MWCNTs, and 96% DLE and 72% DRE for GA-MWCNTs, with pH-controlled drug releaseThe unloaded nanocarrier displayed high biocompatibility, whilst the drug-loaded systems showed slightly higher cytotoxicity than DOX in cancerous cells. Overall, the systems could effectively target cancer cells and deliver DOX to them.[118]
SWCNT-PEG-PEI-FA-CSin vitro: MCF-7 cells74% DLE and approximately 60% DRE, with pH-mediated drug releaseExcellent dispersibility, cellular uptake, and antitumor activity was observed in this system. This nanocarrier caused apoptosis in cancer cells by triggering ROS overproduction.[6]
Fucoidan-decorated silica–carbon nano-onion (FSCNO)-HMin vitro: HUVEC, NCI/ADR-RES, A2780ADR, and OVCAR-8 cells; in vivo: NU/NU nude mice xenografted with NCI/ADR-RES, A2780ADR, and OVCAR-8 cells4% DLC and approximately 60% DRE for DOX, with NIR-triggered drug releaseThe co-delivery of DOX and P-gp pump inhibitor (HM) to counteract chemotherapy resistance increased DOX bioavailability and cytotoxicity. The nanosystem could effectively target and treat both drug- and non-drug-resistant tumour models with decreased side effects. FSCNOs also display photothermal capabilities.[119]
ATRA-NDin vitro: HepG2, MCF-7, and CRL1730; in vivo: HepG2- and MCF-7-induced tumour-bearing BALB/c nude mice25% DLC and >80% DRE, with pH-triggered drug releaseCo-delivery of DOX and ATRA in conjunction with ultrasound treatment. ATRA enhances DOX cytotoxicity, whilst ultrasound enhances nanocarrier permeability into tumour blood vessels. This approach resulted in effective DOX delivery and dramatic tumour growth inhibition.[120]
ND-PEG-HYD-FAin vitro: HeLa, HepG2, MCF-7, and CHO cells; in vivo: BALB/c mice inoculated with HepG2 cells8% DLC, 85% DRE, with pH-triggered drug releaseThis nanocarrier has fluorescence imaging capabilities. Low drug release at pH 7 was due to the use of a cleavable hydrazone linkage. The FA-targeting ligand facilitated endocytosis and the rapid build-up of nanocarrier in cells. In vivo, experiments showed better tumour inhibition along with minimal cardiotoxicity, hepatotoxicity and nephrotoxicity when compared to free DOX.[121]
CNT/HA-DMPEin vitro: MDA-MB-231 and A2780 cells20% DLC, 18% DRE, with pH-triggered drug releaseThis biocompatible nanocarrier is highly stable in biological buffers. The system is easy to prepare and exhibits good targetability towards CD44-overexpressing cancer cells, resulting in a remarkable increase in DOX efficacy.[22]
AL-PEG-MWCNTsin vitro: OPM2, HOS, MCF-7, 3T3-L1, and Raji Burritt’s lymphoma cells; in vivo:
CD-1, Ath/nu, and CB.17 SCID mice; ex vivo: C57BL/6J bones
35% DLC, 51% DREThis system uses individual MWCNTs to improve DOX pharmacokinetics. Its biocompatibility was confirmed through neoplastic transformation, chromosomal aberration, and cytotoxicity assays. Treatment-related weight loss was eliminated in a lymphoma in vivo model, whilst retaining DOX efficacy.[122]
Alginate–urea CDsin vitro: MFC-7 cells>70% DLE, ~45% DRE, with pH-triggered drug releaseThis fluorescent CD system is nontoxic and could be incorporated into hydrogels as a toughening agent. The nanohybrid displayed controlled drug release.[123]
Pharmaceutics 15 01545 i019
Enzalutamide—androgen
receptor antagonist
TP-PEG-Aminated GQDin vitro: C4-2B and LNCaP cells; in vivo: BALB/c nude mice bearing C4-2B tumours60% DLE, 95% DRE, with redox-sensitive drug releaseThis nanocarrier was rapidly uptaken by prostate cancer cells via endocytosis. This caused the effective inhibition of said prostate cancer cell growth in vitro and enhanced targetability and reduced side effects compared to the free drug in vivo.[124]
Pharmaceutics 15 01545 i020
Epirubicin (EPI)—
topoisomerase II inhibitor
SWCNT-DSPE-HAin vitro: A549, AGS, and Taxol-resistant A549 cells47% DLC, 60% DRE, with pH-mediated drug releaseThis system displayed high cancer cell targetability and biosafety. By facilitating the accumulation of EPI in cells by CD44-receptor-mediated endocytosis and preventing EPI efflux by P-gp, this system markedly improved the EPI anticancer efficacy in drug-resistant cancer cells.[125]
Fe3O4-MWCNTsin vitro: T24 and 5637 cells; in vivo: MNU-induced rat-bladder tumour model40% DLE, 100% DRE, with magnetic-field-triggered EPI releaseThis system displayed prolonged retention, enhanced antitumor activity, and enhanced cytotoxicity compared to free EPI. Low systemic toxicity was also seen in vivo. The nanocarrier was also simple and quick to make.[126]
NDin vitro: LT2-MYC cells; in vivo: FVB/N mice with MYC-induced tumours48% DLE, >80% DRE, with pH- and intracellular charged protein-triggered releaseThis nanocarrier prevents the efflux of EPI by ABC transporters to counter chemoresistance in cancer stem cells.[127]
CD-TR-TMin vitro: SJGBM2, CHLA266, CHLA200, and U87 cells75% DRE with no information provided on drug releaseNo reference study was performed with healthy cells. The synergistic anticancer effect of the EPI/TM combination, combined with TR-facilitated targeting, drastically reduces cancer cell viability, even at low concentrations.[128]
HPPH/CPP-PEG-GOin vitro: MG-63 cells; in vivo: osteosarcoma xenograft nude mouse modelApproximately 70% DRE, with no quantitative drug loading information providedThe synergistic effect of HPPH-mediated PDT and EPI chemotherapy allows for control over cancer cell growth. The incorporation of CPP further increases nanocarrier effectiveness by improving cancer cell targeting and internalisation.[129]
Pharmaceutics 15 01545 i021
Erlotinib—tyrosine kinase
inhibitor
GO-PEGin vitro: NPC TW01 cells38% DLC, 98% DRE, with pH-mediated drug releaseThis nanoformulation achieved suppression of nasopharyngeal cancer migration, proliferation, and invasion via several molecular mechanisms.[130]
Carboxylated NDsin vitro: A549, NCI-H460, and NCI-H1975 cells~57% DLE, no quantitative drug release data shownThis nanocarrier caused a decrease in drug-resistant cancer cell viability. Efficient uptake of nanocarrier via clathrin-dependent endocytosis was the key to its effectiveness, and the system was preferentially consumed by cancer cells.[131]
GOin vivo: mice30% DLC, 80% DRE, with pH-triggered drug releaseThe drug was released in a quick-burst fashion.[132]
Pharmaceutics 15 01545 i022
Etoposide (Et)—
topoisomerase II inhibitor
GO-COOHin vitro: HepG2 and RPMI-1640 cells83% DLE, 98% DREThe nanocarrier improved the cytotoxic effect of Et, with no change to its apoptosis pathway. The sustained release of Et offered by this formulation allowed for higher cytotoxicity and efficiency compared to the free drug.[133]
oxiCNH/PEG-PAin vitro: A549 and A549R cells; in vivo: BALB/c nude mice inoculated with A549R cells39% DLC, 81% DRE, with NIR-triggered Et releaseSlow and steady delivery of Et was observed at neutral pH, which was accelerated 1.5× upon NIR irradiation. CNHs are also photosensitizers, and the synergistic effect of PTT and Et chemotherapy killed multidrug-resistant cells by combating P-gp-mediated drug efflux.[134]
FA-CβCDT-MSCDin vitro: HeLa and HepG2 cells14% DLC, 25% DRE, with pH-mediated drug releaseThis nanocarrier displayed preferential targeting of folate-receptor-overexpressing cells. The encapsulation of Et in cyclodextrin prevented premature drug release.[135]
oxiMWCNT-PEG-Asoin vitro: DMS53 and NCIH2135 cells45% DLC, 88% DRE, with pH-sensitive drug releaseThis biocompatible nanocarrier allowed for the cellular internalisation of negatively charged nucleic acids, such as Aso. Aso binding increased the chemosensitivity of drug-resistant lung cancer cells, leading to superior cytostatic efficacy compared to the free drug. This system also displayed good aqueous dispersibility and low haemolytic activity.[136]
Pharmaceutics 15 01545 i023
Gefitinib (GEF)—tyrosine
kinase inhibitor
GO-PVPin vitro: PA-1 and IOSE-364 cells46% DLE, 35% DRE, with GEF release at neutral pHThis biocompatible nanocarrier is a combination therapy with quercetin and was found to enter cells via receptor-mediated endocytosis. The synergistic effect of GEF and quercetin results in significant therapeutic efficacy with ovarian cancer cells, higher than that of drugs delivered separately.[137]
PEG-CQD-PVA-PLAin vitro: NCI–H522 cells~65% DRE, with no quantitative drug loading data shownThe PLA microspheres degrade via a hydrolytic reaction in acidic conditions, releasing GEF. The nanosystem delivered the drug directly to lung cancer cells, resulting in a significant decrease in the IC50 value compared to free GEF.[138]
GO nanosheets 43% DLC, 51% DRE, with pH-triggered drug releaseEffective control over GO nanosheet size was achieved; GEF was converted to nanocrystals and then loaded onto GO.[139]
GO-HAin vitro: A549 and HELF cells; in vivo: BALB/c nude mice bearing A549 tumours13% DLC, 60% DRE, with redox-dependent GET release (glutathione-mediated)This nanocarrier displayed efficient cancer cell uptake via CD44-receptor-mediated endocytosis, resulting in a significant enhancement of the GEF efficacy. The system showed stronger tumour inhibition than the free drug in vivo, with no obvious side effects.[140]
Pharmaceutics 15 01545 i024
Gemcitabine (GEM)—
ribonucleotide reductase
inhibitor
HA-PEG-MWCNTin vitro: HT-29 cells; in vivo: SD rats bearing HT-29 tumours90% DLE, ~85% DRE, with a pH-mediated, sustained drug release profileA reduction in haemolytic activity and remarkable improvement in pharmacokinetic parameters compared to free GEM was observed. This was due to improved cellular internalisation, leading to enhanced anticancer effects.[141]
SWCNT-PEGin vitro: A549 and MIA PaCa-2 cells; in vivo: B6 athymic nude mice (both healthy and with A549 tumours)37% DLC, ~80% DRE, with pH-triggered drug release, via a hydrolysable ester bondThis nanocarrier system accumulates in tumour cells and releases considerable amounts of GEM, resulting in the inhibition of tumour growth and a reduction in GEM side effects. This system also improved the stability of GEM.[142]
PEG-Fe3O4@GO@
mSiO2-FA
in vitro: A431 cells14% DLC, 85% DRE, with pH-triggered drug releaseThis system demonstrated enhanced GEM cytotoxicity and cellular uptake.[143]
ND-PEI-PAA-PEG-GFLGin vitro: BxPC-3; in vivo: BALB/c nude mice xenografted with BxPC-3 tumoursNo quantitative drug loading/release studies were performedSignificant nanocarrier stability was observed in physiological conditions, with long-term circulation due to PEG attachment and enzyme-sensitive GEM release. This system showed similar anticancer effects in vitro and a significant increase in antitumour effects in vivo compared to free GEM.[8]
ND-PEGin vitro: AsPC-1 cellsNo quantitative drug loading/release data were providedThe fluorescent NDs provide imaging and cell-tracking capabilities, and whilst no cytotoxicity enhancement compared to free GEM was seen, the nanocarrier successfully delivered GEM directly to pancreas cancer cells.[144]
ND@PHEA-co-POEGMEAin vitro: AsPC-1 cells7% DLC, ~100% DRE, with pH-triggered drug releaseGEM was incorporated into HEA polymer and then loaded onto NDs. This resulted in slow, sustained GEM release delivered directly to cancer cells, with a similar IC50 value to free GEM.[145]
GO/MMT/CSin vitro: MDA-MB-231 cells99% DRE, with no qualitative drug loading data shown and pH-triggered drug releaseGEM intercalated between MMT silicate layers, preventing burst release. The unloaded nanocarrier is nontoxic, and the sustained release of GEM from the system results in the excellent growth inhibition of breast cancer cells.[146]
FA-CS/Fe3O4/GO 22% DLC, 83% DRE, with pH-triggered drug releaseThis system was tested in simulated cancer fluid and simulated human blood, and it is also responsive to external magnetic fields.[147]
rGOin vitro: A549, HEL-299, and NIH-3 T3 cells; in vivo: BALB/c mice (both healthy and xenografted with A549 tumours)No quantitative drug loading/release data shown. NIR-triggered drug releaseThis system displayed excellent in vitro cytotoxicity against multiple lung cancer cell lines and low systemic toxicity, and a significant enhancement in antitumour activity compared to the free drug in vivo.[148]
CDin vitro: MCF-7 and HeLa cells34% DLC, 54% DRE, with pH-mediated drug releaseThis highly fluorescent nanosystem increased the potency of GEM with minimal cytotoxic effects on healthy cells. This was due to the effective transport and delivery of GEM to cancer cells.[149]
MWCNT-LEin vitro: MCf-7 cells; in vivo: BALB/c nude mice bearing MCf-7 tumoursApprox. 31% DLE, with no qualitative drug release dataThis combination chemotherapy and PTT nanocarrier achieved good antitumour activity in vitro and in vivo, with reduced side effects seen in animal studies.[150]
Pharmaceutics 15 01545 i025
Imatinib—tyrosine kinase
inhibitor
N-prGO-CMC 74% DLC, 58% DRE, with pH-triggered IM releaseThe drug is bound to the nanocarrier via π–π stacking and hydrogen-bonding interactions.[151]
ZnO/CNT@Fe3O4in vitro: CML-derived K562 cellsNo quantitative drug loading/release data shownThe system causes a significant decrease in leukaemia cell viability and metabolism. Cell death was caused by apoptosis via an ROS-dependent mechanism.[152]
GQDin vitro: RPMI 8226, MDA-MB-231, and NCI-ADR/RES cellsNo quantitative drug loading/release data givenThe system demonstrated efficient internalisation and remarkable cytotoxicity for cancer cells and was shown to localise in nuclei of neoplastic cells.[153]
Pharmaceutics 15 01545 i026
Irinotecan—
topoisomerase I inhibitor
Fe3O4-GO-CS-PEG 3% DLC, 90% DRE, with pH-triggered drug releaseA large percentage of drug bound to this nanocarrier was released at neutral pH. This system has magnetic targeting capabilities.[154]
CD-PEG-BTin vitro: MDA-MB231 and MCF-7 cells23% DLC, 90% DRE, with NIR-triggered drug releaseThis combination chemotherapy/PTT nanosystem caused drug-resistant breast cancer cell death via necrosis and apoptosis pathways. The CDs also have fluorescence imaging capabilities.[155]
Fe3O4-GO-CS-UA-GRP-SLP2in vitro: U87 cells; in vivo: BALB/c nude mice with U87 tumours58% DLC, 62% DRE, with pH-responsive drug releaseThis system displayed excellent targeted drug delivery, antitumour efficacy, and prolonged animal survival in brain tumour models using intravenous administration coupled with magnetic guidance. A 4.9× increase in drug uptake compared to the free drug was measured in vitro. A 6.5× enhancement in the ability to cross the blood–brain barrier compared to the free drug was seen in vivo. Highly biocompatibility was also seen in vivo.[156]
Pharmaceutics 15 01545 i027
Lobaplatin—DNA alkylating agent
E2-PEG-CNTin vitro: MCF-7 cells; in vivo: healthy C57BL/6 miceNo qualitative drug loading/release data shownThis nanocarrier has sustained release properties, with no obvious side effects and an increase in retention time compared to the free drug in vivo. The effective killing of cancer cells was seen in vitro due to E2-mediated targeting.[157]
PEG-CNT-FITCin vitro: HepG2 cells72% DLC, 80% DRE, with pH-triggered, sustained drug release and fluorescence imaging propertiesThis nanocarrier can effectively enter and kill liver cancer cells, with increased cytotoxicity observed up to 72 h.[158]
Pharmaceutics 15 01545 i028
Methotrexate (METX)—
nucleotide synthesis inhibitor
oxiSWNH-PEG-Tfin vitro: MAD-MB-231 and HepG2 cells; in vivo: ICR mice carrying H22 tumours15% DLC, 55% DRE; this system utilises pH-triggered drug releaseThe drug is released quickly at neutral pH, which could cause toxicity due to premature leakage. This system displayed favourable tumour targeting, cytotoxicity, and accumulation.[25]
CMC-GOin vitro: NIH-3T3 and HT-29 cells; in vivo: BALB/c mice and nude mice xenografted with HT-29 tumours39% DLC, 82% DRE, with pH-triggered drug releaseThis system reduced drug toxicity against healthy cells and facilitated a higher plasma concentration, superior tumour cytotoxicity, and liver cancer metastasis inhibition compared to free METX.[159]
Hydroxylated C60 fullerenein vitro: MDA-MB-231 cells; in vivo: Wistar ratsNo quantitative drug loading data shown, with 85% DRE and pH-sensitive drug releaseThis nanosystem drastically increased plasma half-life and AUC compared to the free drug, resulting in a large reduction in its IC50 value. Enhanced bioavailability, erythrocyte compatibility, protein binding, and haemolysis in human blood compared to free METX were also observed.[23]
AF-FA-99mTc-MWCNTin vitro: A549 and MCF 7 cells; in vivo: New Zealand rabbits and FR+ EAT-bearing mice33% DLC, >85% DRE, with pH-triggered drug release achieved via a cleavable ester linkageEffective targeting and treatment of folate-receptor-overexpressing cancer cells with reduced side effects and increased efficacy in vivo was observed. This nanocarrier also had fluorescence imaging and radio-tracing capabilities.[160]
Pharmaceutics 15 01545 i029
Mitomycin C (MMC)—
DNA alkylating agent
TAT-graphenein vitro: OCM-1 and ARPE-19 cells22% DLC, 45% DRE, with pH-triggered drug release; however, the release in acidic and neutral environments was very similarThis system could specifically target cancer cells over healthy cells in a co-culture environment. The nanocarrier localised in the cancer cell nuclei, resulting in strong growth suppression.[161]
CDin vitro: MCF-7 cellsApproximately 80% DRE, with no quantitative drug loading information provided and pH-mediated MMC releaseMMC was bound to the CDs via hydrogen bonding. This nanocarrier showed high affinity towards cancer cell membranes and could effectively enter them and accumulate. This resulted in a significant improvement in anticancer potency compared to free MMC.[162]
SWCNT-PEG-CWKG(KWKG)6in vitro: A549 cellsApproximately 80% DRE, with pH-mediated MMC releaseThe unloaded nanocarrier showed high biocompatibility whilst the drug-loaded system exhibited similar anticancer efficacy to free MMC.[163]
Graphene-BODIPY-PEGin vitro: HeLa cells10% DLC, with no quantitative drug release data givenThis nanocarrier possesses excellent photothermal conversion efficiency and ROS production capabilities for combination PTT/PDT. The system also has fluorescence and photothermal imaging capabilities and displayed outstanding anticancer effects.[14]
Pharmaceutics 15 01545 i030
Mitoxantrone (MTX)—topoisomerase II inhibitor
GO-Fe2O3-MitPin vitro: A549 cells19% DLC, 38% DRE, with magnetic-field-triggered MTX releaseMitP grafting improves the drug loading capability of this nanocarrier. Successful targeting and disruption of tumour mitochondria was achieved, causing cell death.[164]
oxiMWCNTin vitro: NIH3T3 and MDA 231 cells95% DLC, 30% DREMTX was bound to oxiMWCNTs via electrostatic interactions. This formulation resulted in increased MTX efficacy; however, the system was more toxic to healthy cells than cancerous cells.[165]
NDin vitro: MDA-MB-231 and MDA-MB231-ABCG2 cells87% DLE, 80% DRE, with pH- and soluble-protein-triggered MTX releaseMTX release was found to be higher in FBS than in water, suggesting that the presence of soluble biological matter increases the DRE. A marked increase in MTX retention and efficacy was observed when using this nanocarrier.[166]
oxiSWCNT-PEG-FAin vitro: HeLa cells~35% DLE, 55% DRE, with pH-mediated, sustained drug releaseThis system selectively targeted cancer cells.[167]
EXO-GO-CO-γPGAin vitro: MDA-MB-231 and BEAS-2B cells73% DLE, 56% DRE, with pH-mediated, sustained drug releaseThis nanocarrier displays excellent cancer cell targetability. The attachment of exosomes was found to improve drug loading, pH response, and biocompatibility.[168]
rGO-PEG-SBin vitro: 4T1, CT26, and bone marrow macrophages + DCs harvested from BALB/c mice; in vivo: BALB/c mice with 4T1 tumours48% DLE, with no quantitative drug release data provided and NIR-triggered drug releaseThe synergistic combination of PTT, chemotherapy, and immunotherapy facilitated the destruction of local primary tumours and distant metastases in an in vivo model. rGO acted as a photosensitizer, whilst the SB immunotherapeutic increased effectiveness of rGO and MTX by TGF-β inhibition.[169]
Pharmaceutics 15 01545 i031
Oxaliplatin (OP)—
DNA alkylating agent
GO-PNVCL-PGAin vitro: MCF-7 cells12% DLC, 80% DRE, with pH- and thermal-responsive OP releaseImproved cytotoxicity compared to free OP against breast cancer cells was observed. The blank nanocarrier is nontoxic.[170]
GO-HSA NPsin vitro: HFFF261% DLE, ~97% DRE, with pH-mediated, sustained drug releaseThe use of HSA nanoparticles increased nanocarrier biocompatibility.[171]
MWCNT-PEGin vitro: HT29 cells43% DLC, no quantitative drug release dataA drastic increase in cytotoxicity towards human bowel cancer cells was observed.[172]
GO-CB [7]/Ce6/AQ4N/ADA-HAin vitro: B16 and L02 cells; in vivo: C57BL/6 mice (both healthy and carrying B16 tumours)58% DRE, with spermine-triggered OX releaseSignificant antitumour efficacy was observed, resulting from synergistic PTT (GO)/PDT (Ce6), and chemotherapy (OP and AQ4N) properties. The enhanced hypoxia resulting from PTT bolsters the effects of chemotherapy drugs. This strategy has the benefit of being able to noncovalently attach nonaromatic drug molecules via host–guest complex formation.[173]
TAT-BT-PEI-MWCNTin vitro: C6 glioma (cell and tumour spheroid) and CHEM-5 and L02 cells; in vivo: mice bearing C6 tumours19% DLC, with no quantitative drug release data shownThe nanocarrier shows enhanced blood–brain barrier penetration compared to free OX, resulting in a significant decrease in the IC50 value. The system shows low cytotoxicity towards healthy cells; however, a build-up of cerebrospinal fluid was noticed during treatment of in vivo models.[174]
GO-CS-FAin vitro: LX-2 and SKOV3 cells34% DLC, ~80% DREThis nanoformulation shows similar potency to free OX in ovarian cancer cells and good biocompatibility.[48]
CDin vitro: L929, HeLa, and HepG2 cells4% DLC, with redox-sensitive drug releaseThe CDs have multicoloured emission capabilities and high fluorescence stability. This system shows good biocompatibility, bio-imaging, and anticancer effects both in vitro and in vivo.[175]
Pharmaceutics 15 01545 i032
Paclitaxel (PTX)—microtubule growth inhibitor
GO-MA/FAin vitro: MDA-MB-231 cells; in vivo: SD rats with DMBA-induced mammary carcinoma40% DLC, 71% DREPTX was attached to GO via π–π stacking and hydrophobic interactions. This system inhibited cancer cell growth in vitro and reduced tumour size in vivo via cell cycle arrest and apoptosis. Highly specific targeting and drug release for folate-receptor-overexpressing cancers was observed.[176]
SWCNT/DOA-PEG-FAin vitro: MCF-7 cells; in vivo: athymic nude mice with MFF-7-induced tumours and BALB/c miceNo quantitative drug loading or release data were providedThis system displayed high specificity, biocompatibility, and efficacy compared to free PTX in vitro. In vivo, studies revealed significant tumour growth inhibition, with no side effects on the blood and major organs of mice observed.[177]
PLA composite nanofibers/C70 fullerenein vitro: HepG2 cellsNo quantitative drug loading information, 72% DREControl of in vitro PTX release profile was achieved by varying fullerene content. Successful control of cancer cell growth was achieved using this nanoformulation.[178]
Rf-MWCNTsin vitro: MCF-7 cells; in vivo: SD rats82% DLE, 99% DREHaemolysis in human blood was much lower than free PTX. This system also showed better cytotoxicity than free PTX, with low systemic toxicity, favourable biodistribution, and renal excretion observed in vivo.[179]
Au–N-doped carbon nanotube cup (NCNC)in vitro: MDSC cells; in vivo: B16 melanoma cells inoculated into C57BL/6 mice36% DLE, with no quantitative drug release data shownThe PTX-containing NCNCs capped with Au nanoparticles exhibited strong surface-enhanced Raman scattering effects, allowing for extremely sensitive detection. A single injection of nanocarrier solution given to an in vivo model significantly reduced tumour growth and eliminated tumours in ~30% of mice. This system targeted lymphoid tissues surrounding tumours to boost the host immune system response.[180]
Graphene-PLA-PEGin vitro: U-138 cells; in vivo: athymic nude Foxn1nu mice implanted with U-138 tumours4% DLC, 6% DREThis nanocarrier is nontoxic and expressed long-term sustained drug release over 19 days. The system was found to be six times more potent than free PTX in vitro and is capable of reducing U-138 cell viability despite low drug loading and release.[181]
HMCSin vitro: CNE cells; in vivo: nude mice with CNE tumours19% DLC, no release experiments were carried out with PTXHMCSs themselves exhibited strong antitumour effects through PTT.[86]
CNT-PMAA self-assembled micellesin vitro: L929 and HeLa cells36% DLE, 74% DRE, with pH-triggered drug releaseThe blank nanocarrier is nontoxic, with low haemolysis observed. Higher anticancer activity than free PTX was also noted. The self-assembly of CNT-PMAA is pH-dependant, and at low pH, the nanocarrier disassembles.[182]
FA-CD-GOxin vitro: MDA-MB-468 (cells and tumour spheroids); TNBC and HEK293 cells82% DLE, 95% DRE, with GOx and PTX release occurring at pH 7.4GOx-induced cancer starvation, which had a synergistic effect with PTX chemotherapy, and resulted in significant cancer cell death. This biocompatible nanocarrier also efficiently targets cancer cells over normal cells.[13]
FCDbin vitro: NIH3T3 and B16F10 cells82% DLC, with no quantitative release data givenThis nanocarrier targeted cancer cells in a co-culture with healthy cells due to its biotin-targeting ligand. It caused the selective sensing and activation of H2O2, and it also has fluorescence imaging capabilities.[183]
Pharmaceutics 15 01545 i033
Sorafenib—tyrosine kinase
inhibitor
rGOin vitro: SGC7901 cellsNo quantitative drug loading/release dataA significant increase in cytotoxicity compared to the free drug in gastric cancer cells was observed, with apoptosis being the main mechanism of cell death.[184]
CS nanoparticles-FAin vitro: HepG2, HDFa, and HT29 cells19% DLC, 91% DRE, with pH-triggered, sustained drug release profileThis system showed negligible toxicity to healthy cells and enhanced anticancer properties compared to the free drug.[185]
CNT-PEGin vitro: HepG2 cells; in vivo: Wistar rats, both healthy and with DENA-induced liver tumours95% DLC, ~100% DREA negligible change in potency or morphology was observed over a 3-month stability study. This system displayed superior anticancer abilities compared to the free drug both in vitro and in vivo.[186]
DSPE-PEG-NDin vivo: healthy rats and BALB/c mice carrying BGC-823 tumours~90% DRE, with no quantitative drug loading data shownA 14-fold increase in drug concentration in tumour tissues compared to the free drug was observed. This caused considerable growth inhibition for the in vivo gastric cancer model. A 7.64× increase in oral bioavailability compared to the free drug was also seen in vivo.[187]
Pharmaceutics 15 01545 i034
Tamoxifen (TAM)—oestrogen receptor antagonist
rGOin vitro: MCF-7 cells; in vivo: BALB/c mice bearing MCF-7 tumours This system displayed lower cytotoxicity than free TAM but had promising photothermal properties. No quantitative drug loading/release data were shown. The nanocarrier has excellent stability.[188]
NGR-SWCNT-PF68in vitro: 4T1 cells; in vivo: BALB/c mice carrying 4T1 tumours This targeted nanocarrier could effectively enter cancer cells whilst retaining TAM cytotoxicity. Receptor-mediated tumour targeting combined with PTT capabilities resulted in significant anticancer efficacy in vivo. No quantitative drug loading/release data shown.[189]
MWCNT-LEin vitro: MCF-7 cells28% DLC, no quantitative drug release dataEnhanced cellular uptake, apoptosis, and antitumour activity (compared to free TAM) were observed. This system also has PTT properties. LEN acts as both a dispersant and a potent anticancer agent itself. Combination chemo–PTT results in cancer cell destruction at low drug concentrations.[190]
C60 fullerene-Glyin vitro: MCF-7 cells; in vivo: Wistar rats66% DLC, 85% DRE, with pH-triggered TAM releaseThis nanocarrier vastly improved the pharmacokinetic properties and cytotoxicity of TAM against breast cancer cells. The system displayed minimal haemolysis towards human blood. The bioavailability, half-life, and cancer cell penetration of the drug were all markedly improved.[191]
TEG-MWCNT-quercetinin vitro: MDA-MB-231 cells; in vivo: Wistar ratsNo quantitative drug loading info, 93% DRE, and pH-triggered drug release using a TAM prodrugThis haem-compatible formulation reduced the IC50 values and increased the uptake in drug-resistant cells. These favourable properties carried over to in vivo studies, resulting in enhanced efficacy, pharmacokinetics, and biocompatibility.[192]
DES-graphenein vitro: MCF-7 cellsNo quantitative drug loading/release dataNanocarrier possesses acute, selective anticancer activity achieved through intracellular ROS-production-triggering cell cycle arrest. [193]
Pharmaceutics 15 01545 i035
Temozolomide—DNA
alkylating agent
GO-Fe3O4in vitro: rat glioma C6 cells90% DLC and 74% DRE, with pH-mediated drug releaseThe blank nanocarrier is biocompatible, whilst the loaded system showed better inhibitory effects than the free drug in rat glioma cells. This formulation also has strong magnetic properties.[194]
rGOin vitro: LN229 cells84% DLC and 83% DRE, with electrochemically controlled drug releaseThis system retained the anticancer potency of temozolomide.[195]
Pharmaceutics 15 01545 i036
Topotecan—topisomerase I
inhibitor
GO/TT/CisP/cholesterol-DOX encapsulated in a DSPE-PEG nanocellin vitro: HeLa cells29% DLC and 40% DRE for TT, with pH-triggered drug releaseThis nanocarrier has fluorescence imaging capabilities and displays remarkably higher anticancer efficacy compared to the free-drug combination due to the synchronised targeting of multiple targets in cancer cells at once.[196]
PEI-GO-TT-CisPin vitro: HeLa cells51% DLC for TT, no quantitative drug release dataThis nanocarrier has subcellular-organelle-targeting capabilities and can effectively impair the mitochondria of cervical cancer cells, leading to cell death. This resulted in a 4.4× decrease in IC50 compared to the free-drug cocktail.[197]
TT/GO-CD/DOXin vitro: HeLa cells16% DLE and 77% DRE for TT, with sustained, pH-triggered drug releaseThis system demonstrated superior anticancer efficacy compared to free drugs and single-drug-loaded nanocarrier due to the synergistic effect between the drugs and GO.[198]
Pharmaceutics 15 01545 i037
Vinblastine—mitosis inhibitor
CQDin vitro: Hela, HGC-27, A549, MCF-7, CF-STTG, and Vero cells; in vivo: NOD-SCID mice carrying A549 tumours 95% DLC, with no qualitative drug release dataThe cytotoxicity of vinblastine was reduced in normal cells and increased in cancer cells compared to the free drug. Significant inhibition of tumour growth was observed in vivo, with no liver toxicity. The synergistic combination of chemotherapy and PTT allowed for the control of cancer cell growth.[199]
Pharmaceutics 15 01545 i038
Vincristine—mitosis inhibitor
oxiSWCNH-PEG-mAbin vitro: MCF-7 and HUVEC cells; in vivo: ICR mice carrying H22 tumours35% DLC, 80% DRE, with pH-triggered, sustained drug releaseEffective cancer cell killing in MCF-7 cells and treatment of liver cancer in mice with superior efficacy compared to the free drug. The system displayed reduced systemic toxicity compared to the free drug due to mAb-facilitated targeting.[200]
Table 2. Index of anticancer drug molecules in the CNM nanocarrier database.
Table 2. Index of anticancer drug molecules in the CNM nanocarrier database.
DrugPageDrugPage
5-fluorouracil5epirubicin23
6-mercaptopurine6erlotinib24
anastrozole7etoposide24
bortezomib7gefitinib25
capecitabine7gemcitabine26
carboplatin7imatinib27
carmustine9irinotecan28
chlorambucil9lobaplatin29
cisplatin9methotrexate29
curcumin13mitomycin C30
cyclophosphamide15mitoxantrone30
cytarabine15oxaliplatin31
dabrafenib16paclitaxel33
dasatinib16sorafenib34
daunorubicin16tamoxifen35
decitabine17temozolomide36
docetaxel17topotecan36
doxorubicin19vinblastine37
enzalutamide23vincristine37
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Mohan, H.; Fagan, A.; Giordani, S. Carbon Nanomaterials (CNMs) in Cancer Therapy: A Database of CNM-Based Nanocarrier Systems. Pharmaceutics 2023, 15, 1545. https://doi.org/10.3390/pharmaceutics15051545

AMA Style

Mohan H, Fagan A, Giordani S. Carbon Nanomaterials (CNMs) in Cancer Therapy: A Database of CNM-Based Nanocarrier Systems. Pharmaceutics. 2023; 15(5):1545. https://doi.org/10.3390/pharmaceutics15051545

Chicago/Turabian Style

Mohan, Hugh, Andrew Fagan, and Silvia Giordani. 2023. "Carbon Nanomaterials (CNMs) in Cancer Therapy: A Database of CNM-Based Nanocarrier Systems" Pharmaceutics 15, no. 5: 1545. https://doi.org/10.3390/pharmaceutics15051545

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop