Next Article in Journal
The TGF-β Family in Glioblastoma
Next Article in Special Issue
Pancreatic Neuroendocrine Tumors: Signaling Pathways and Epigenetic Regulation
Previous Article in Journal
Whole-Exome Screening and Analysis of Signaling Pathways in Multiple Endocrine Neoplasia Type 1 Patients with Different Outcomes: Insights into Cellular Mechanisms and Possible Functional Implications
Previous Article in Special Issue
A Cancer-Specific Monoclonal Antibody against Podocalyxin Exerted Antitumor Activities in Pancreatic Cancer Xenografts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Graphene Oxide Nanoparticles and Organoids: A Prospective Advanced Model for Pancreatic Cancer Research

by
Shaoshan Mai
and
Iwona Inkielewicz-Stepniak
*
Department of Pharmaceutical Pathophysiology, Faculty of Pharmacy, Medical University of Gdańsk, 80-210 Gdańsk, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2024, 25(2), 1066; https://doi.org/10.3390/ijms25021066
Submission received: 3 December 2023 / Revised: 2 January 2024 / Accepted: 4 January 2024 / Published: 15 January 2024
(This article belongs to the Special Issue Therapeutic Targets in Pancreatic Cancer)

Abstract

:
Pancreatic cancer, notorious for its grim 10% five-year survival rate, poses significant clinical challenges, largely due to late-stage diagnosis and limited therapeutic options. This review delves into the generation of organoids, including those derived from resected tissues, biopsies, pluripotent stem cells, and adult stem cells, as well as the advancements in 3D printing. It explores the complexities of the tumor microenvironment, emphasizing culture media, the integration of non-neoplastic cells, and angiogenesis. Additionally, the review examines the multifaceted properties of graphene oxide (GO), such as its mechanical, thermal, electrical, chemical, and optical attributes, and their implications in cancer diagnostics and therapeutics. GO’s unique properties facilitate its interaction with tumors, allowing targeted drug delivery and enhanced imaging for early detection and treatment. The integration of GO with 3D cultured organoid systems, particularly in pancreatic cancer research, is critically analyzed, highlighting current limitations and future potential. This innovative approach has the promise to transform personalized medicine, improve drug screening efficiency, and aid biomarker discovery in this aggressive disease. Through this review, we offer a balanced perspective on the advancements and future prospects in pancreatic cancer research, harnessing the potential of organoids and GO.

1. Introduction

Pancreatic ductal adenocarcinoma (PDAC) is one of the most aggressive and deadliest forms of cancer, known for its rapid progression and poor prognosis. Despite significant advancements in cancer research and treatment modalities, PDAC continues to have a dismal five-year survival rate of less than 11%, with this figure dropping to less than 3% in patients with advanced stages of the disease [1]. This high mortality rate is attributed to several factors including the cancer’s high malignancy, insidious onset, lack of distinct symptoms, challenging anatomical location, low resection rate, and high recurrence rate.
Common signs and symptoms of PDAC, often appearing only in advanced stages, include jaundice, abdominal and back pain, unexplained weight loss, and digestive difficulties. Globally, pancreatic cancer is the seventh leading cause of cancer-related deaths, with its incidence and mortality rates closely aligning due to its aggressive nature. According to the World Health Organization, there were over 450,000 new cases and 430,000 deaths worldwide in 2020, reflecting its substantial impact on global health [2,3].
Currently, standard treatment modalities for PDAC include surgery, chemotherapy, and radiation therapy. While surgical resection offers the best chance for a cure, only a small fraction of patients is eligible for this option. The majority of pancreatic cancer cases are diagnosed at an advanced stage, rendering surgery infeasible [4]. The mainstay of treatment for advanced pancreatic cancer, chemotherapy, has limited efficacy. Drugs such as gemcitabine have been the standard, but they offer only modest improvements in survival and are frequently associated with significant side effects. In addition, as it is a broadly used drug, almost all pancreatic cancer patients eventually develop resistance to this drug [5]. FOLFIRINOX (oxaliplatin, irinotecan, leucovorin, 5-fluorouracil) could directly improve the overall survival (OS) rate of patients with metastatic pancreatic tumor (HR 0.76, 95% Cl 0.67–0.86, p < 0.001) but had no benefit on progression-free survival (PFS) [6]. Standard radiotherapy options, which typically deliver 40 to 60 Gy in 1.8–2.0 Gy fractions, offer minimal to no survival benefit for patients with locally advanced unresectable pancreatic cancer (LAPC) who have undergone chemotherapy [7]. Its role is limited due to the proximity of the pancreas to critical organs, which increases the risk of damage to surrounding tissues. Therefore, these limitations of treatment underscore the urgent need for innovative therapeutic strategies.
In the realm of research, organoid models have emerged as a promising tool. These organoids accurately represent the genomic, proteomic, and morphological characteristics of parental tumors, providing a more faithful replication of patient tumors compared to traditional in vitro 2D cell cultures [8]. Organoids are advantageous for their ability to be passaged long-term, cryopreserved, and genetically manipulated without significant genomic or epigenetic alterations, making them ideal for constructing living biobanks and advancing personalized therapy approaches. These organoid models have been instrumental in understanding disease progression, studying tumor microenvironment interactions, and contributing to precision medicine [9,10].
Graphene oxide (GO), a derivative of graphene, stands out in the realm of nanomaterials due to its unique properties and potential applications, particularly in the biomedical field. GO is characterized by its chemical and mechanical stability, biocompatibility, and a two-dimensional structure that allows for extensive surface modification [11]. These modifications can be tailored with various functional groups such as epoxide, hydroxyl, and carboxyl, enabling the attachment of biomolecules such as proteins, DNA, and RNA [12]. This adaptability makes GO an attractive candidate for a range of applications, including drug and gene delivery, phototherapy, and bioimaging. In medical research, GO’s ability to be used in drug delivery systems, diagnostics, tissue engineering, and gene transfection is particularly noteworthy [13]. Its solubility in water and intrinsic fluorescence properties in the visible/near-infrared spectrum enhance its suitability for these applications [14,15]. Recent studies have also explored the possibility of using GO for specific targeting of cancer cells, potentially opening new avenues for cancer therapy, including the inhibition of metastasis.
In the context of tumor therapy, GO is often used as a nano-carrier due to its ability to load a large number of hydrophobic drugs containing benzene rings. Its nano-network structure and hydrophobicity play a crucial role in this process [16,17]. Additionally, GO’s surface properties can be affected by pH changes [18]. It remains stable at a neutral pH but becomes less stable at more extreme pH values [19]. This characteristic is particularly useful in cancer therapy, as tumor tissues generally have a more acidic microenvironment compared to normal tissues. Therefore, at lower pH values, such as those found in tumor tissues, the protonation weakens the hydrogen bond interaction between the drug and GO. This pH-sensitive property can be incorporated into the design of an anticancer drug delivery system, making GO an intelligent nano-carrier that releases drugs more effectively in the acidic environment of tumor tissues [20]. In addition, it is often functionalized with various components, such as synthetic polymers such as polyethylene glycol (PEG) or natural polymers [21]. This functionalization can be achieved through covalent modification, which may alter the original structure of GO, or non-covalent methods, which do not affect its native structure. These modifications enhance the stability of GO in physiological solutions and facilitate its use in drug delivery [22].
In this review, we aim to provide a comprehensive overview of the recent developments in organoids (Figure 1) and GO, focusing on their roles in restructuring the cancer microenvironment, their advancements, and potential biomedical applications. We consolidate current knowledge on the properties and applications of organoids and GO (Figure 2) in cancer research. This includes discussing the ongoing challenges in these fields and exploring the potential synergistic future of organoids and GO platforms. Our primary focus was on their application in personalized treatment strategies for pancreatic cancer, emphasizing the need to address current limitations to fully realize their potential in this context.

2. Organoid Models for Studying Cancer

2.1. Origin of Organoids

2.1.1. Organoids Generation from Resected Tissues

Over the last decade, most origins of cancer organoids have been derived from primary tumors and metastatic lesions. They are collected from surgical resections. Sato et al. described the generation and expansion of patient-derived organoids from colon tissue [23]. Pancreatic cancer organoids were built to reveal genes and pathways altered during disease progression [24]. Lung cancer organoids recapitulated the tissue architecture of the primary lung tumors [25]. To date, these cancer organoids include colorectal [26,27], ovarian [28,29,30], prostate [31,32], pancreatic [33,34], liver [35,36], bladder [37,38], lung [39,40,41], gastric [42,43], brain [44,45], esophagus [46,47], and endometrium [48,49,50,51].

2.1.2. Organoids Generation from Biopsy

The one of main challenges of organoids is tissue accessibility. Except for small fragments of surgically resected tissue, solid and liquid biopsies are used as sources of tumor organoid cultures. For example, bladder cancer organoid lines can be established efficiently from patient biopsies acquired before and after disease recurrence [52]. Human prostate cancer organoids were retrieved from biopsy specimens and circulating tumor cells successfully in long-term cultures [53]. Protocols for establishing organoids from human ovarian cancer biopsies have been developed [54,55]. In addition, colorectal organoids can be derived from metastatic biopsy specimens with a high success rate and genetically represent the metastasis they were derived from [56]. Circulating tumor cells (CTCs) are a rare subset of cells found in the blood of patients with solid tumors; the CTC-derived pre-clinical model matches the mutation with the primary tumor, which could be applied in clinical research for the evaluation of disease progress [57].

2.1.3. Organoid Generation from Pluripotent Stem Cells

Organoids developed from human pluripotent stem cells (PSCs) are particularly useful for tissue shortage, such as the nervous system and retina [58,59]. Human-PSC-derived organoids with components of all three germ layers (ectoderm, mesoderm, endoderm) have been generated, resulting in the establishment of a new human model system. hPSCs, including human induced pluripotent stem cells (hiPSCs) and human embryonic stem cells (hESCs). Both of them share an unlimited proliferative ability and the developmental potential to generate all three germ layers [60]. Colonic organoids derived from hiPSC for modeling colorectal cancer provide an efficient strategy for drug testing [61]. Patient-derived iPSCs from a Li-Fraumeni syndrome (LFS) family were used to investigate the role of mutant p53 in the development of osteosarcoma (OS) [62]. iPSC-derived embryoid bodies of a patient with hereditary c-met-mutated papillary renal cell carcinoma (PRCC) were generated spontaneous aggregates organizing in structures that expressed kidney markers such as PODXL and Six2 [63]. iPSCs were differentiated into pancreatic ductal and acinar organoids that recapitulate the properties of the neonatal exocrine pancreas, as well as the mutated KRAS G21D oncogene in organoids to form pancreatic cancer [64]. Engineering prostate cancer from iPSC is used to develop preclinical tools in prostate cancer studies [65].

2.1.4. Organoids Generation from Organ-Specific Adult Stem Cells

Adult stem cell (ASC)-derived organoids were shown to be genetically stable over long periods [66]. The genetic modification of small intestine, stomach, liver, and pancreas organoids has opened up avenues for the manipulation of ASC in vitro, which could facilitate the study of human biology and allow gene correction for regenerative medicine purposes [67]. Human intestinal stem cells mutated via CRISPR/Cas9 technology for colorectal cancer organoid cultures remain genetically and phenotypically stable [68]. Efficient protocols for the culture of breast cancer (BC) organoids that are established from mammary epithelium serve as a representative collection of well-characterized BC organoids for BC research and drug discovery [69]. On exposure to pregnancy signals, endometrial organoids derived from ASC develop the characteristics of early pregnancy [70].

2.1.5. Organoid Generation from 3D Printing

The lack of precise architectures and size are the key limitations of tissuee- and PSC-derived organoids. 3D bioprinting of organoids has shown its advantages in recent years. It is a promising technology to precisely position the biological materials, living cells, and growth factors for the computer-aided generation of bioengineered structures. Bioinks and bioprinters are two necessary and key factors used in 3D printing. Natural polymer-based bioinks such as agarose [71], alginate [72], collagen [73], hyaluronic acid [74], gelatin, and matrigel [75] are used for printing. Bioprinting methods include inkjet printing (nozzle-based), extrusion bioprinting (nozzle-based), light-based bioprinting stereolothography, digital light processing (nozzle-free), and laser-assisted bioprinting (nozzle-free).
It is difficult to restructure pathological cancer tissue in organoids and standardize the number and size of organoids in current organoid culturing. However, 3D printing is a prospective tool for cancer organoid treatment. Bioprinting can reconstitute the tumor microenvironment, for example, control matrix properties and integrate vascular networks. Glioblastoma-on-a-chip is a 3D bioprinting model in which patient-derived glioblastoma cells are co-cultured with endothelial cells to form a cancer–stroma concentric ring structure, which serves the purpose of screening for effective treatment modalities for patients [76]. In a 3D bioprinting collagen model, biomimetic chimeric organoids were developed by co-bioprinting breast cancer cells and normal mammary epithelial cells, which showed a more significant increase in 5-hydroxymethylcytosine in the chimeric structure than the tumoroid one [77]. A 3D bioprinting organotypic pancreatic adenocarcinoma model was used to restructure a cancer microenvironment that consisted of endothelial and pancreatic stellate cells [78]. A high-throughput bioprinting system was used to construct a Matrigel substrate via co-bioprinting ovarian cancer cells (OVCAR-5) and fibroblasts (MRC-5) [79]. Hela cells and gelatin/alginate/fibrinogen hydrogels have been used to construct cervical tumor models in vitro via extrusion-based bioprinting [80]. A hepatic cancer model was fabricated through a two-step stereolithographic bioprinting method [81].

2.2. Extracellular Matrix of Organoids

2.2.1. Animal-Derived Matrix

The extracellular matrix (ECM) is a dynamic polymer network used to influence the cell biology of neoplastic and tumor microenvironments (TMEs). Hydrogels derived from the decellularized basement membrane of murine Engelbreth-Holm-Swarm (EHS) sarcoma, such as Matrigel or Geltrex, are the most commonly used matrix for organoid culturing [82,83]. EHS matrices have been widely adopted in cancer organoids, relying on their potential to recapitulate the 3D tumor structure and provide growth factors that maintain TME. For human cancer organoid culturing, however, the EHS matrix is limited by its ill-defined, poorly tunable animal-derived scaffolds and batch-to-batch variability [84,85]. It is associated with increased collagen deposition and remodeling in tumor progression. Collagen type I matrices are also a less expensive alternative for cancer organoids. Similarly, animal sources that face the same problem as EHS matrices, and their microstructures (for example, fibril diameter, and alignment) are highly dependent on the rate of pH and temperature changes during gelation [86,87].

2.2.2. Engineered Matrix

Engineered matrices have yet to be routinely applied to human cancer organoid cultures; however, compared with animal-derived matrices, they offer high batch-to- batch reproducibility, standardization of cancer organoid formation, mimic the specific composition and structure of the native ECM, and high-throughput screening. Primary human glioblastoma (GBM) organoids are encapsulated in a hybrid material comprising synthetic polyethylene glycol (PEG) decorated with the RGD integrin-binding peptide and crosslinked with recombinant hyaluronic acid (HA) [88]. PEG-based dynamic hydrogels functionalized with the basement membrane protein laminin enable reproducible intestinal stem cell (ISC) expansion and organoid formation [89].
Engineered matrices for organoid cultures can optimize specific organoid systems, including stiffness, matrix viscoelasticity, and degradability. Organoids based on purified silk protein and alginate polysaccharides enable ECM systems to retain their superior homogeneity and reproducibility [90,91]. Fibrin matrices provide suitable physical support and naturally occurring Arg-Gly-Asp (RGD) adhesion domains on the scaffold, as well as supplementation with laminin-111; these are key parameters required for robust organoid formation and expansion [92]. A hybrid matrix comprising HA and elastin-like protein (ELP) regulate the growth rate of intestinal organoids and their formation efficiency [93]. A synthetic hydrogel extracellular matrix was designed for pancreatic organoids and replicated the phenotypic traits characteristic of the tumor environment in vitro [94].

2.3. Tumor Microenvironment of Organoids

2.3.1. Culture Medium

Organoid media are necessary components in the establishment of cancer microenvironments in vitro. For example, Wnt 3a, R-spondin, Noggin, and epidermal growth factor are not dispensable for organoid growth. To preserve organoid heterogeneity, the omponents of the organoid media must be very complex. Wnt protein is at the heart of organoid technology. According to researchers who identify and develop the driver of TME, intestinal organoids are strictly dependent on Wnt ligands for survival and growth [95]. A set of growth factors are critical components of organoid media, which include R-spondins and BMP signaling antagonists such as Noggin or Gremlin 1 [96]. A B27 supplement comprises various nutrients, which increase cell survival rate, promote the formation of tumor spheres, and prevent sphere adhesion during lung cancer organoid culturing [97]. N-acetylcysteine is an antioxidant directly scavenging ROS and partially via ERK1/2 activation, and also a precursor of antioxidant glutathione (GSH) [98]. FGF7 and FGF10 are mainly used to induce lung organoid branching [99]. EGF belongs to a family of growth factors that drive proliferation in different organoids, such as colon cancer organoids [23], breast cancer organoids [69], pancreatic cancer organoids [24], and prostate cancer organoids [53]. A83-01 is a TGF-β/activin receptor-like kinase 5 (ALK5) inhibitor. The overexpression of TGF-β induces epithelial-to-mesenchymal transition (EMT) and facilitates immunosuppression, ECM deposition, and angiogenesis. Therefore, A83-01 inhibits the EMT of cancer organoids by inhibiting TGF-β [26]. In addition, other factors play different roles in organoid culturing, such as the N2 supplement, nicotinamide, prostaglandin E2, and gastrin 1, which induce differentiation, promote cell survival, and so on [100].

2.3.2. Non-Neoplastic Cells

The co-culturing of primary tumor epithelia with endogenous, syngeneic tumor-infiltrating lymphocytes (TILs), which comprise patient-derived organoids (PDOs), enables the successful modeling of the immune checkpoint blockade (ICB) with anti-PD-1- and/or anti-PD-L1, expanding and activating tumor antigen-specific TILs and eliciting tumor cytotoxicity [101]. Co-culturing with cancer-associated fibroblasts (CAFs) enhances the organoid-forming ability of CD44+ cells, as well as the expression of CD44 and OCT-4 at the protein level [102]. Multi-cell type organotypic co-culture models include stromal cells, immune cellular components, and pancreatic cancer organoids. The activation of myofibroblast-like cancer-associated fibroblasts and tumor-dependent lymphocyte infiltration were observed in these models [103]. The microenvironment includes CAFs, tumor endothelial cells (TECs), tumor-associated macrophages (TAMs), and regulatory T cells (Tregs), which influence and recapitulate cancer progression [104,105].

2.3.3. Angiogenesis

Primary HUVECs were co-cultured with primary colorectal tumoroids to observe induced angiogenesis via tube formation assay. The assay evidenced that tube formation increased in a dose-dependent manner upon treatment with the pro-angiogenic factor vascular endothelial growth factor A(VEGF-A) [106]. Co-cultures of hepatocellular carcinoma (HCC) cell lines or patient-derived xenograft organoids with endothelial cells exhibited the upregulation of MCP-1, IL-8, and CXCL16, influencing tumor necrosis factor (TNF) and angiocrine signaling, which generate an inflammatory microenvironment by recruiting immune cells [107]. Interactions with vasculature were found in spheroids and organoids upon 7 days of co-culture with space of Disse-like architecture in between hepatocytes and endothelium, resulting in a stable, perfusable vascular network [108]. Vascularized spheroids were generated from a non-adherent microwell culture system of human umbilical vein endothelial cells, human dermal fibroblasts, and human umbilical-cord-blood-derived mesenchymal stem cells to develop and assemble vascular spheroids with cerebral organoids [109]. The microvascular network cultured with patient-derived tumor organoids presented highly angiogenic features. ECM components and the culture media composition were adjusted to coculture patient-derived colon organoids, which form a self-assembled microvasculature under intravascular perfusion by using a microfluidic platform [110]. The relations between platelets and cancer organoids are also prospective for tumor angiogenesis studies [111,112].

3. A Graphene Oxide Platform for Cancer Research

3.1. Properties of Graphene Oxide

Graphene, a two-dimensional honeycomb lattice of carbon atoms, is typically synthesized via chemical vapor deposition or mechanical exfoliation of graphite, noted for its exceptional electrical, mechanical, and thermal properties [113]. GO created using the Hummers method introduces oxygen-containing functional groups to graphite, resulting in a material with a large surface area and functional versatility, albeit with reduced electrical conductivity compared with graphene [114,115]. Reduced graphene oxide (r-GO), obtained by removing these oxygen groups from GO, restores some of graphene’s electrical and structural features [116]. Graphene quantum dots (GQDs), small graphene fragments synthesized through top–down or bottom–up methods, exhibit unique optical and electronic properties due to quantum confinement and edge effects [117]. Both GO and r-GO, rich in functional groups, are adaptable for diverse modifications and applications, particularly in biosensors and environmental remediation, while GQDs find use in fluorescence-based sensors and electronics [118]. Graphene’s integration into composites boosts their mechanical, thermal, and electrical characteristics, with r-GO also being leveraged in similar applications for its balance between conductivity and functional compatibility [119]. We discuss the detailed properties of GO in the following sections.

3.1.1. Mechanical Properties

GO consists of stacked layers of graphene sheets that are held together by oxygen-containing functional groups, creating a layered, lamellar structure. The structure of graphene is a monolayer of two-dimensional (2D) one-atom-thick planar, sp2-hybridized carbon arranged in six-atom rings [120]. GO is a complex material due to its amorphous and non-stoichiometric atomic composition. Currently, there are no precise analytical techniques available for the thorough characterization of GO materials and their analogs. According to previous research, several structural models have been proposed, including models by Hofmann, Ruess, Scholz-Boehm, Nakajima-Matsuo, Lerf-Klinowski, and Dekany. In terms of regular lattices, they consist of discrete repeated units, and the Lerf-Klinowski model is currently considered the most widely accepted configuration [121,122]. Brittle fracture ensues when the material is subjected to a critical stress, typically around 130 GPa, in accordance with its intrinsic strength. Because of its high values of E (elastic modulus) and σint (intrinsic strength), graphene is regarded as an exceptionally robust material for structural applications. Graphene oxide paper, on average, exhibits a fracture strength of 80 MPa and an elastic modulus of 32 GPa [123].

3.1.2. Water Dispersibility

GO is highly hydrophilic and readily disperses in water and other polar solvents due to the presence of hydrophilic functional groups. This property is advantageous for various applications, such as nanocomposites and biomedical applications. The presence of oxygen-containing functional groups such as -OH, -COOH, and epoxide groups on the surface makes GO hydrophilic. Félix Mouhat et al. evidenced that GO is chemically active in water, acquiring an average negative charge of the order of 10 mCm−2 [124]. The hydrophilicity of graphene oxide with different particle sizes and pH values was characterized by the water contact angle. And Xuebing Hu et al. found that the water contact angle of the different graphene oxides decreased from 61.8° to 11.6°, which indicates graphene oxide has excellent hydrophilicity [125].

3.1.3. Thermal Properties

Graphene stands out as an exemplary thermal conductor, demonstrating an impressive thermal conductivity range of 2000–5000 W/mK [126]. Nevertheless, the introduction of oxygen functional groups onto the surface of graphene oxide (GO) disrupts the lattice symmetry and induces localized strain, resulting in a substantial reduction in thermal conductivity by orders of magnitude (2–3 orders of magnitude, to be precise) [127]. A remedy for this attenuation in thermal conductivity involves the partial reduction of GO through chemical reactions with reducing agents, notably hydrazine and its derivatives. This process culminates in the creation of reduced graphene oxide (rGO) by effectively extracting oxygen functional groups. An interesting observation has been made regarding the thermal conductivity of GO, which exhibits a continuous decrease as the degree of oxidation escalates [128]. Studies show the successful formation of GO/RGO–protein complexes with enhancement in structural/thermal stability due to various interactions at the nano–bio interface and their utilization in various functional applications [129].

3.1.4. Electrical Properties

Graphene oxide is inherently non-conductive, necessitating the removal of a significant portion of its oxygen groups for conversion into reduced graphene oxide (rGO) to enhance its electrical conductivity. Researchers have observed a distinct difference in electric conductivity between R-I-Ph-GO/PI films and R-GO/PI films, primarily attributed to the formation of a sp2-hybrid carbon network within the graphene oxide structure [130]. Notably, the electrical properties of GO films exhibit sensitivity to both humidity levels and applied voltage amplitude [131]. At low humidity, GO films demonstrate poor conductivity, akin to insulators. Conversely, under high humidity conditions, GO film conductivity markedly increases because of enhanced ion conduction mechanisms, offering insights into tailored electrical properties for GO-based materials in applications influenced by environmental factors, particularly humidity.

3.1.5. Chemical Properties

GO offers superior dispersibility in various mediums, including water, diverse solid matrices, and organic solvents. This property makes it highly versatile. GO can be effectively combined with polymer or ceramic matrices to form composite materials, often resulting in improved electrical and mechanical properties [121]. However, GO does come with certain limitations, such as the potential for agglomeration or overlapping of GO sheets. Nonetheless, the thin and flat structure of GO sheets allows for flexibility in making structural and morphological modifications. This flexibility is further enhanced by the presence of oxygen functional groups in GO, which serve both as sites for functionalization and as spacers for molecular absorption. These attributes facilitate the incorporation of GO into various nano composites and nano-morphologies. By employing structural modifications and functionalization through covalent and noncovalent bonding interactions, GO finds applications in a wide array of real-world uses, including filtration membranes [132], electrochemical sensors [133], hydrogen storage devices [134], battery electrodes [135], supercapacitors [136,137], and microjet engines [138].

3.1.6. Optical Properties

GO has a range of remarkable optical properties driven by its electronic configuration. These properties include structure-dependent absorption and Raman spectra, which provide insights into its chemical composition and the extent of functionalization-induced disorder [139]. In contrast to pristine graphene, GO displays photoluminescence across a spectrum encompassing ultraviolet, visible, and near-infrared wavelengths, contingent upon its structural variations. Reduced graphene oxide (rGO), on the other hand, exhibits the capacity to absorb radiation across an extensive wavelength range spanning from ultraviolet to terahertz frequencies [140]. To investigate the effects of modifications on fluorescence behavior, various agents, such as polyethylene glycol (PEG) polymers, metal nanoparticles (including Au and Fe3O4), and folic acid (FA) molecules, have been employed to functionalize the surface of GO [141]. rGO was functionalized with L-arginine (L-Arg) that on the optically active support generated an effective optical chemosensor for the determination of Cd (II), Co (II), Pb (II), and Cu (II) [133]. The marriage between integrated photonics and GO has led to the birth of integrated GO photonics, which has become a very active and fast-growing branch of on-chip integration of 2D materials in order to achieve novel functionality of integrated photonic devices [140].

3.1.7. PH-Sensitivity

Graphene oxide-based nanomaterials can be affected by pH changes in their surface properties. The tumor environment is generally more acidic (pH 6.4~7) than normal cells (pH 7.4) [142,143]. A dual-targeting drug delivery and pH-sensitive controlled release system GO–Fe3O4 nanohybrid has been established [144]. β-cyclodextrin grafted L-phenylalanine functionalized graphene oxide is a versatile nanocarrier for pH-sensitive doxorubicin delivery [145]. GO was functionalized covalently with pH-sensitive poly(2-(diethylamino) ethyl methacrylate) (PDEA) by surface-initiated in situ atom transfer radical polymerization. Simple physisorption by π-π stacking and hydrophobic interactions on GO-PDEA can be used to load camptothecin (CPT), a water-insoluble cancer drug that is released only at lower pH levels normally found in a tumor environment but not in basic and neutral pH circumstances [146]. The constructed graphene oxide hybrid cyclodextrin-based supramolecular hydrogels could respond to NIR light, temperature, and pH, which could be beneficial for the controlled release of cargoes. Graphene oxide sheets not only acted as a core material to provide additional cross-linking but also absorbed NIR light and converted NIR light into heat to trigger the –sol–gel transition [147].

3.2. Graphene Oxide in Cancer Diagnosis

The most effective approach for curbing the advancement of cancer is the development of innovative diagnostic tools that enable the detection of the disease at its early stages. The early detection of carcinoma cells and the continuous monitoring of their activities hold significant importance in the realms of clinical diagnostics, toxicity monitoring, and safeguarding public health. Detecting and monitoring tumor cells during their early stages play a pivotal role in preventing the progression of cancer. This is particularly crucial in cases such as pancreatic cancer, which is challenging to diagnose at an early stage and often leaves limited possibilities for rescuing patients in advanced stages of the disease. Biomedical imaging technologies serve as highly efficient tools for tumor diagnosis, providing invaluable insights that can effectively guide tumor therapeutics. GO serves as a versatile agent for bioimaging functions due to its unique physical and chemical properties (Figure 2).

3.2.1. Magnetic Resonance Imaging

Magnetic resonance imaging (MRI) stands out as a non-invasive and non-ionizing diagnostic method, renowned for its exceptional spatial and temporal resolution. In the context of MRI, nanomaterials based on GO functionalized with paramagnetic metals have demonstrated great promise. Here, dendrimers featuring amino group caps (DEN) are skillfully grafted onto GO nanosheets which lateral sizes in the range of 40–380 nm (mean size ∼175 nm). This grafting process serves as a crucial step, enabling the subsequent functionalization of GO with gadolinium diethylene triamine pentaacetate (Gd-DTPA) and prostate stem cell antigen (PSCA) monoclonal antibody (mAb). Remarkably, the in vivo results obtained from magnetic resonance imaging validate the utility of GO-DEN(Gd-DTPA)-mAb as a targeted contrast agent for prostate tumor imaging [148]. Scientists have integrated GO with superparamagnetic iron oxide nanoparticles (Fe3O4 NPs). These Fe3O4 NPs serve a dual purpose, functioning as both biocompatible magnetic drug delivery enhancers and magnetic resonance contrast agents for MRI. The synthesized GO-Fe3O4 conjugates exhibit an average size of 260 nm and demonstrate low cytotoxicity levels, which are on par with those observed in GO alone [149]. Researchers focused on the synthesis of hybrid nanocomposites, specifically graphene oxide–zinc ferrite (GO-ZnFe2O4), which are further conjugated with doxorubicin (DOX) for applications in cancer therapy and MRI-based diagnosis. GO-ZnFe2O4 and GO-ZnFe2O4/DOX ranging from 5 to 100 μg/mL were investigated, and the key components are zinc ferrite (ZnFe2O4) nanoparticles (NPs), which serve as MR imaging contrast agents [150].
The size of magnetic nanoparticles (MNPs) in graphene oxide (GO) composites critically influences their performance, with smaller MNPs enhancing surface reactivity, ensuring superparamagnetic behavior, and improving biological penetration and distribution, while also affecting stability and toxicity profiles [151]. This size-dependent variation in physical and magnetic properties is fundamental in tailoring GO-MNP conjugates for specific applications, particularly in biomedical fields such as MRI [152].

3.2.2. Fluorescence Imaging

Fluorescence lifetime imaging (FLIM) is a non-invasive technique reliant on photons emitted by fluorescent probes, frequently employed for monitoring pathological tissue. A graphene oxide–MnO2–fluorescein (GO–MnO2–FL) nanocomposite was synthesized and applied for the detection of reduced glutathione (GSH). GSH, an essential endogenous antioxidant, plays a pivotal role in cellular defense against reactive oxygen species (ROS), thereby maintaining cellular activity. Notably, the GO–MnO2–FL (100 μg/mL) nanocomposite exhibited selective imaging of cancer cells, owing to the higher GSH content observed in cancer cells as compared with normal cells [153]. A non-invasive and targeted technology for early diagnosis of oral squamous cell carcinoma (OSCC) involves the synthesis of nano-graphene oxide (NGO) nanoparticles. These nanoparticles are designed with specificity to the gastrin-releasing peptide receptor (GRPR), achieved through the incorporation of GRPR-specific peptides AF750-6Ahx-Sta-BBN. This approach enables precise and non-invasive near-infrared fluorescence imaging targeted at OSCC [154]. GO-based fluorescent DNA nanomaterials offer a promising avenue for the in vitro diagnosis and therapy of liver tumor cells [155].

3.2.3. Photoacoustic Imaging

Photoacoustic Imaging (PAI) stands as a robust diagnostic tool hinging on the photoacoustic (PA) effect. The distinctive capability of PAI lies in its ability to furnish optical absorption contrast and achieve high-resolution imaging, rendering it particularly well-suited for applications involving deep tissue and organ imaging [156]. PA imaging is a noninvasive imaging modality that depends on the light absorption coefficient of the imaged tissue and the injected PA-imaging contrast agents. PA imaging integrates the excellent contrast achieved in optical biomedical imaging with the deep penetrability of ultrasound (US) imaging. Thus, PA imaging can be used for the imaging of deeper tissue compared to other optical imaging technologies [157].
A nanotheranostic agent has been fabricated by direct deposition of Bi2Se3 nanoparticles on graphene oxide (GO) in the presence of polyvinylpyrrolidone (PVP) using a one-pot solvothermal method. GO/Bi2Se3/PVP nanocomposites (2 mg/mL) could serve as an efficient bimodal contrast agent to simultaneously enhance X-ray computed tomography imaging and photoacoustic imaging in vitro [158]. The reduced graphene oxide coated gold nanorods (r-GO-AuNRs) and highly efficient heat transfer process through the reduced graphene oxide layer, r-GO-AuNRs (0.125 mg/mL), exhibit excellent photothermal stability and significantly higher photoacoustic amplitudes; therefore, r-GO-AuNRs can be a useful imaging probe for highly sensitive photoacoustic images [159]. A sandwich-type gold nanoparticle coated reduced graphene oxide (rGO-AuNP) as an effective nanotheranostic platform for the second near-infrared (NIR-II) window photoacoustic (PA) imaging-guided photothermal therapy (PTT) in ovarian cancer [160]. With nanoparticles composed of a liquid gallium core with a reduced graphene oxide (RGO) shell (Ga@RGO) of tunable thickness, the high near-infrared absorption of RGO results in a photothermal energy conversion of light to heat of 42.4%. This efficient photothermal conversion, combined with the large intrinsic thermal expansion coefficient of liquid gallium, allows the particles to be used for photoacoustic imaging, that is, the conversion of light into vibrations that are useful for imaging [161].
Indocyanine green (ICG)-loaded, polyethylene glycosylated (PEG), reduced nano-graphene oxide nanocomposite (rNGO-PEG/ICG) is a new type of fluorescence and photoacoustic dual-modality imaging contrast. The nanocomposite is demonstrated to possess greater stability, longer blood circulation time, and superior passive tumor targeting capability, which can be a promising candidate for further translational studies on both the early diagnosis and image-guided therapy/surgery of cancer [162].

3.2.4. Raman Imaging

Raman scattering (SERS) is widely used due to its non-invasiveness, ultrasensitivity, and high spatial resolution. Since molecular vibrations are strongly related to the molecular structure, condition, and environment, the combination of spontaneous Raman scattering can be used to monitor both the location and condition of biological molecules in living cells [163]. GO possesses characteristic fingerprints in Raman spectra; therefore, it is also used for Raman imaging with Au and Ag nanoparticles loaded as SERS substrates [164]. GO/gold nanoparticle (AuNP) hybrids with folic acid (FA) binding are prepared as a multifunctional platform in bioimaging. FA is the targeting agent, AuNPs work as surface-enhanced Raman scattering substrates, and GO takes the role of both supporting the AuNPs with FA and acting as a Raman probe [165]. Methylene blue-loaded mesoporous silica-coated gold nanorods on graphene oxide (MB-GNR@mSiO2-GO) (36 µg/mL) were developed as an all-in-one photo-nanotheranostic agent for intracellular surface-enhanced Raman scattering (SERS) imaging-guided photothermal therapy (PTT)/photodynamic therapy (PDT) for cancer [166].
Afua A. Antwi-Boasiako and colleagues reported the use of bioconjugated 2D graphene oxide (bio-GO) nanostructures, with the average lateral size of the layered GO sheets being approximately 0.08–0.1 μm. These were used as probes for breast cancer cells (SKBR3), demonstrating excellent discrimination over other types of circulating tumor cells by monitoring the ‘turn-off’ of the Raman signal [167]. Lin Yang et al. designed silver nanoparticles deposited on graphene oxide for ultrasensitive surface-enhanced Raman scattering immunoassay of cancer biomarkers, which made the detection of prostate-specific antigen (PSA) serum samples from prostate cancer patients satisfactory and thereby demonstrating that sensitive enzyme-assisted dissolved AgNPs SERS immunoassays of PSA have potential applications in clinical diagnosis [168].

3.2.5. Computed Tomograph

Computed tomography (CT) is a widely adopted disease diagnosis method in clinical settings because of its non-invasiveness and high spatial resolution properties, which are based on its high atomic number and X-ray absorption. Graphene oxide/gold nanorod (GO/GNR) nanohybrids were synthesized with a GO- and gold-seed-mediated in situ growth method. Upon injection of the GO/GNRs (50 μg/mL) into xenograft tumors, excellent CT imaging properties and photothermal effect were obtained [169]. Zhan Li et al. evidenced that Ag nanoparticles (AgNPs) are composited on the surface of GO to promote its X-ray absorption, and then simvastatin is coinjected in mice of renal dysfunction to eliminate in vivo toxicity-induced by AgNPs [170]. One new composite contrast agent based on Ln and graphene matrices was developed for multi-energy computed tomography [171].
Graphene nanoplatelets (GNPs), synthesized via potassium permanganate-based oxidation and exfoliation followed by reduction with hydroiodic acid (rGNP–HI), intercalated with manganese ions within the graphene sheets, and covalently functionalized with iodine, exhibit excellent potential as bimodal contrast agents for magnetic resonance imaging (MRI) and CT [172]. A novel system for synergistic cancer therapy was developed based on bismuth sulfide (Bi2S3) nanoparticle-decorated graphene functionalized with polyvinylpyrrolidone (PVP) (named PVP-rGO/Bi2S3). GO nanosheets with an average diameter of ~100 nm and a thickness of ~1.2 nm were prepared via a modified Hummer’s method. Due to the obvious NIR and X-ray absorption ability, the PVP-rGO/Bi2S3 nanocomposite could be employed as a dual-modal contrast agent for both photoacoustic tomography and X-ray computed tomography imaging [173]. By using a solvothermal method in the presence of polyethylene glycol (PEG), BaGdF5 nanoparticles are firmly attached to the surface of GO nanosheets to form the GO/BaGdF5/PEG nanocomposites, thus enabling effective dual-modality MR and X-ray computed tomography (CT) imaging of the tumor model in vivo and indicating the potential applications of dual-modality MR/CT imaging of cancers [174].

3.3. Graphene Oxide in Cancer Treatment

3.3.1. Delivery System

Graphene-based materials have high specific surface areas and low toxicity. GO-based nanomaterials have been extensively studied in cancer treatment (Figure 2), for instance, as a new type of nanocarriers to deliver various therapeutic anticancer agents. In this section, we discuss recent advancements in GO delivery systems for cancer research. Notably, we compile a list of significant recent studies on functionalized GO delivery systems (refer to Table 1). This list showcases the progress and innovations in the application of GO for targeted delivery systems of cancer therapy.

Drug Delivery

GO can be functionalized with polymers. Chitosan (CS) is an amino polysaccharide, a hydrophilic, biocompatible, non-toxic, and biodegradable polymer of glucosamine and acetylglucosamine [175]. To synthesis GO-CS is involved in amide coupling between the COOH group on the GO and the amino group of CS. CS improves the solubility of GO sheets in acidic media. Moreover, GO-CS results in changes in particle size and zeta potential as a function of pH [176,177]. Yan et al. prepared GO-CS and investigated its potential as a nanoadjuvant. GO-CS (50 μg/mL) significantly activated RAW264.7 cells and stimulated more cytokines for mediating cellular immune responses [178].
PEG is a polymer of repeating ethylene ether units that is widely used for several pharmaceutical and biomedical applications. It is soluble in aqueous and organic media, being defined as biocompatible, biodegradable, non-toxic, non-immunogenic, and is classified as “Generally Regarded as Safe” (GRAS) by the FDA [179]. Nano-graphene oxide (NGO) was synthesized its biological applications were explored in order to develop functionalization chemistry to impart solubility and compatibility to NGO in buffer solutions and other biological media by covalently grafting PEG star polymers onto its chemically activated surfaces and edges [180]. To evaluate it as a potential anti-metastatic agent, GO was modified with polyethylene glycol to form PEG-modified GO (PEG-GO). PEG-GO did not show apparent effects on the viability of breast cancer cells (MDA-MB-231, MDA-MB-436, and SK-BR-3) or non-cancerous cells (MCF-10A), but inhibited cancer cell migration in vitro and in vivo. An analysis of cellular energy metabolism revealed that PEG-GO significantly impaired mitochondrial oxidative phosphorylation (OXPHOS) in breast cancer cells but not in non-cancerous cells [181]. A de novo drug delivery nanosystem (~128 nm) based on gold nanoparticles (GNPs), decorated PEG, and folate (FA)-conjugated GO was designed to load with doxorubicin hydrochloride (DOX) as a model anticancer drug. Its drug-loading capacity as well as pH-dependent drug release behavior were investigated [182].
Hyaluronic acid (HA) is a naturally mucopolysaccharide, biocompatible, non-immunogenic, and biodegradable polysaccharide, consisting of alternating units of D-glucuronic acid and N-acetyl-D-glucosamine [183]. Numerous tumor cells overexpress several receptors that have a high binding affinity for HA, while these receptors are poorly expressed in normal body cells. Graphene quantum dots (GQD) were used as drug carriers, and HA was decorated on the surface of GQD to target cancer cells. At the same time, curcumin (CUR) was used as a drug model and loaded on the synthesized nanocarriers. GQD-HA-CUR reduces HeLa cell viability significantly because of the mediation of HA–CD44 for drug cell uptake [184]. Metformin was loaded upon GO via drop-wise addition of 2 mL (10 mg/mL) of metformin into 20 mL of GO dispersion (5 mg/mL). HA was grafted onto metformin-loaded GO nanoparticles as a CD44-targeted anti-cancer therapy for triple-negative breast cancer, which exhibited anti-cancer efficacy at a much lower dosage as compared with metformin alone [185]. Doxorubicin (Dox) and paclitaxel (Ptx) were successfully loaded onto GO-HA that covalently attached HA onto GO. The GO-HA-Dox/Ptx system was significantly better than the GO-Dox/Ptx system at specifically killing CD44-expressing MDA-MB-231 cells but not BT-474 cells without the expression of CD44 [186].
Polyvinyl alcohol (PVA) is a water-soluble polymer synthesized via hydrolysis and radical polymerization of vinylacetate [187,188]. Curcumin was loaded in PVA-sodium alginate/3D-GO hydrogels for studying in vitro drug delivery systems [189]. Magnetic magnesium iron oxide nanoparticles were synthesized via the coprecipitation chemical method and then composited with graphene oxide and modified by polyvinyl alcohol. Paclitaxel (PTX) and docetaxel (DTX) were loaded in the modified magnetic nanocomposites. The generally sustained and controlled release profile of DTX (or PTX) facilitates the application of modified nanocomposite for the delivery of anticancer drugs [190].
Polyacrylic acid (PAA) is a synthetic polymer of acrylic acid monomers, which is biocompatible, non-toxic, pH-sensitive, and mucoadhesive. In aqueous solutions, PAA has an anionic nature because of its carboxylic groups [191]. Nanocomposite systems, consisting of a reduced graphene oxide/polyacrylic acid as a nanocarrier, were integrated with a folic-acid-targeting agent and further modified by Deferrioxamine-M (M: Mn2+ or Gd3+) as the diagnostic MRI contrast agent or Temozolomide as the therapeutic agent. Release studies at a biological of pH 7.4 revealed good stability for TMZ immobilized on the GNs@PAA-FOA/TMZ nanocarrier [192]. Gemcitabine (GEM)—PAA—GO are developed in an explicit solvent medium at two different pH values, which can control drug biodistribution in response to changes in pH that are markers of the tumor environment [193].
Polyvinylpyrrolidone (PVP) is a water-soluble, non-ionic, non-toxic polymer surfactant. The coating of PVP is reported to improve the dispersibility and biocompatibility of GO in physiological buffers. Injectable hydrogel polymeric nanoparticles of PVP cross-linked with N, N′-methylene bis-acrylamide and encapsulating water-soluble macromolecules FITC–dextran (FITC–Dex) have been prepared in the aqueous cores of reverse micellar droplets, which serve as a potential carrier for hydrophilic drugs [194]. A stimuli-responsive polyvinylpyrrolidone-NIPPAm-lysine graphene oxide nano-hybrid was designed and fabricated for the delivery of chemotherapeutic agent fluorouracil (FU) to MCF7 breast cancer cells [195]. Nanocarriers comprising gelatin (G)-PVP coated GO were prepared and loaded with quercetin (QC). Additionally, a dual nanoemulsion water/oil/water with bitter almond oil was developed as a membrane around the nanocomposite to control further drug release. The pH-sensitive drug delivery system showed an 87.5% encapsulation efficiency and a 45% drug loading, which are among the highest values reported up to date. MTT assay and flow cytometry methods revealed a rate of cancer cell death of 53.14%, which was 36.51% in the apoptotic phase [196].
Dextran (Dex) is a hydrophilic natural polymer and a polysaccharide synthesized from the condensation of glucose. Cellular experiments uncover that DEX coating on GO offers remarkably reduced cell toxicity [197]. The non-covalent functionalization of GO with chitosan (CS) and Dex was successfully developed via a layer-by-layer self-assembly technique for anti-cancer drug delivery application. The CS/Dex functionalized GO nanocomposites (GO-CS/Dex) exhibited a diameter of about 300 nm and a thickness of 60 nm and showed a strong cytotoxicity to the cancer cells [198].
Graphene oxide/cationic polyethyleneimine/poly anionic dextran sulfate (GO/PEI/DS) was synthesized via a layer-by-layer self-assembly technique for transdermal anti-cancer drug delivery. DOX was loaded onto folic-acid-conjugated GO and methotrexate (MTX) was loaded onto dextran sulfate (DS). The results revealed that the synthesized dual drug-loaded material showed a good pH-dependent controlled and sustained release profile for both DOX and MTX via the transdermal route of administration in comparison with oral delivery [199]. GO-PEI complexes were loaded with a miR-214 inhibitor which efficiently inhibited cellular miR-214, resulting in a decrease in oral squamous cell carcinoma (OSCC) cell invasion and migration and an increase in cell apoptosis [200].

Gene Delivery

Gene therapy needs a vector that can protect genes from nuclease degradation and facilitate gene uptake with high transfection efficiency. Compared with viral vectors, non-viral gene carriers can avoid immune responses, toxicity, chromosomal integration, and so on. GO-based vectors for gene delivery have been considered as superior vehicles because of their good biocompatibility, biosafety, large surface area, adsorption capacity, and negative charges [201].
Polyethyleneimine (PEI) is a water-soluble cationic polymer with a large number of amino groups. GO was modified with PEG, PEI, and FA for the targeted delivery of small-interfering RNA (siRNA) that inhibits ovarian cancer cell growth [202,203]. Lactosylated chitosan oligosaccharide (LCO)-functionalized GO (GO-LCO) containing quaternary ammonium groups (GO-LCO+) was prepared to realize the hepatocyte-specific targeted delivery of anti-tumor drugs and genes. The GO-LCO+ could be used to load Dox with the loading efficiency of 477 μg/mg and fluorescein FAM-labeled DNA with 4 μmol/g, respectively. Results suggest that the functionalized GO can be used as a nanocarrier for the hepatocyte-targeted co-delivery of anti-tumor drugs and genes with low cytotoxicity, indicating its potential future applications in anticancer drug-and-gene combined therapy [204].
Glycyrrhetinic acid (GA) is employed as a liver-targeting ligand to construct GA, polyethylene glycol (PEG), polyamidoamine dendrimer (Dendrimer), and nano-graphene oxide (NGO) conjugates (GA-PEG-NGO-Dendrimer, GPND) for siRNA delivery. The GPND/siRNA nanocomplex has high safety, targeting, and transfection as well as a prolonged half-life [205]. A modified GO nanocarrier for the co-delivery of siRNA and DOX was designed for enhanced cancer therapy. GO-poly-l-lysine hydrobromide/folic acid (GPF)/DOX/siRNA exhibited gene silencing and tumor inhibition [206]. GO/3-aminopropyltriethoxysilane (APTES) was modified via spermine (GOAS), which acts as a gene delivery system to help the transfection of pEGFP-p53 into breast cancer cell lines [207]. Nanoparticles comprising GO/cationic lipid (GOCL) condense and stabilize plasmid DNA for transfection into human cervical cancer (HeLa) cells [208].

Antibody Delivery

A pH-responsive charge-reversal polyelectrolyte and integrin αⅤβ3 mono-antibody functionalized GO complex was developed as a nanocarrier for the targeted delivery and controlled release of DOX into cancer cells [143]. A GO platform was functionalized with magnetic nanoparticles and a monoclonal antibody specific to the carbonic anhydrase marker. CA IX is a cell surface hypoxia-inducible enzyme functionally involved in adaptation to acidosis that is expressed in aggressive tumors [209]. Multifunctional mesoporous silica nanoparticles (MSNs) have internal fluorescent conjugates and external polydopamine and GO layers. Monoclonal antibody (anti-human epidermal growth factor receptor)-conjugated MSNs showed a higher specificity, which resulted in more enhanced anticancer effects in vitro [210].
GO was modified with a non-toxicity cationic material (chitosan) and a tumor-specific monoclonal antibody (anti-EpCAM) for the delivery of survivin-siRNA (GCE/siRNA). It was demonstrated that GCE/siRNA had a strong antitumor effect in vitro [211]. An antibody-modified reduced graphene oxide (rGO) film efficiently captured circulating tumor cells (CTCs) and minimized the background of white blood cells without complex microfluidic operations [212]. The noncovalent association of anti-HER2 antibody trastuzumab (TRA) with GO generates stable TRA/GO complexes that are capable of rapidly killing osteosarcoma (OS) cells [213].
A single layer of carboxymethylcellulose (CMC) and poly N-vinylpyrrolidone (PVP) was cross-linked through a disulfide bond and deposited on graphene oxide nanoparticles (GO NPs). The NPs were functionalized via monoclonal antibody FA, which showed a high inhibition of Saos2 and MCF7 cell lines in vitro [214]. Yang et al. evidenced the efficient targeting of breast cancer metastasis in an experimental murine model featuring GO conjugated with a monoclonal antibody (mAb) against follicle-stimulating hormone receptor (FSHR), a highly selective tumor vasculature marker in both primary and metastatic tumors [215].
Table 1. Delivery systems in recent studies on functionalized GO.
Table 1. Delivery systems in recent studies on functionalized GO.
Type of NanomaterialsAdvantagesLimitationsCarriers
(Drugs/Gene/Antibody)
Reference
GO-CS-(FA)smaller size, positive surface charge, increased compatibility, high loading capacitythermally unstabledrug: Chitosan/Folic acid [176]
GO–PEG–FA/GNPs–DOXhigh drug loading capacity, acceptable biocompatibility, pH-responsive drug release profileslow drug releasedrug: Doxorubicin hydrochloride[182]
HA-GO-Metselectively targeting CD44+ triple-negative breast cancer (TNBC) cells, anti-cancer efficacy in vitro and in vivocomplex synthetic methoddrug: Metformin[185]
PVA-SA/3D-GOlower swelling and higher releasecomplex synthetic method, percentages of GO influence deliverydrug: Curcumin[189]
(DTX/PTX)-MgFe2O4-GO-PVAsustained and controlled releasecomplex synthetic method without an in vivo testdrug: Paclitaxel, Docetaxel[190]
FU-GO/NHsFU-GO/NHs more cytotoxic than free FU, fast uptake, and temperature and pH-responsive no normal cells for comparison drug: 5-fluorouracil [195]
GO/PEI/DS-DOX-MTXdual drug loading, good selectivity, prolonged drug existence in the blood circulation system, and pH-dependent controlled and sustained releasecomplex synthetic methoddrug: Doxorubicin, Methotrexate[199]
PEG-GO-PEI-FA/siRNAsiRNA condensation and stablecomplex synthetic method without an in vivo testgene: siRNA[203]
GO-LCO+(-FAM-DNA)co-delivery of anti-tumor drugs and genes with low cytotoxicitycomplex synthetic methoddrug: Doxorubicin chloride
gene: DNA
[204]
GA-PEG-NGO-Dendrimer/anti-VEGFa siRNAstability, low toxicity, negligible hemolytic activity, high transfection efficiency, and targetedcomplex synthetic methodgene: anti-VEGFa siRNA[205]
APTES-GO-(pEFGP-p53)Ninety percent of the cells were transfectedunstable during the heating process, complex synthetic method, and no in vivo testgene: EFGP-p53 plasmid[207]
GOCL/DNAhigh transfection, low toxicity, a library of 9 cationic lipids screeningno biological milieu testgene: DNA[208]
GO-Abs/PEI/PAH-Cit/DOXtargeted cancer cells, release by mild acidic pH stimulationcomplex synthetic methodantibody: integrin αⅤβ3 drug: Doxorubicin[143]
GO-MNps-EDC-MAbtargeted drug delivery, enhanced biocompatibilitypotential cytotoxicity, stability, and uniformity of the compositesantibody: CA IX cDNA[209]
MSNs-CPT@A-F@PDA@GOMSNs exhibited stimuli (pH, NIR irradiation)-responsive controlled release, a higher specificity, and efficient cytotoxicity toward cancer cellscomplex synthetic method, without an in vivo testantibody: anti-human EGFR
drug: Cisplatin
[210]
GO-CS-anti-EpCAM-siRNAantitumor effect, accumulates siRNA in tumor tissues, and biosafety carriercomplex synthetic method, no significant differences between GCE/siRNA and Lipo/siRNA for downregulation rates of survivin-mRNA antibody: anti-EpCAM gene: survivin-siRNA[211]
CMC/PVP GO-FA-Curcuminenhanced antiangiogenesis, apoptosis, and tumor growth inhibitioncomplex synthetic methodantibody: folic acid antibody
drug: curcumin
[214]
NOTA-GO-FSHR-mAb stability and high specificity for FSHR, use for early metastasis detection, and targeted delivery of therapeuticscomplex synthetic methodantibody: anti-follicle-stimulating hormone receptor drug: Doxorubicin[215]

3.3.2. Phototherapy

Phototherapy, including photothermal therapy (PTT) and photodynamic therapy (PDT), is a newly developed and encouraging therapeutic strategy, which employs near-infrared (NIR) laser photoabsorbers to generate heat for thermal ablation of cancer cells upon NIR laser irradiation [216].

Photothermal Therapy

PTT operates by transforming radiant light energy into localized heat through external NIR laser irradiation. This process induces hyperthermia, raising the temperature at the tumor site and subsequently causing damage and apoptosis of tumor cells. Owing to its unique advantages of high specificity and minimal invasiveness, PTT has been evidenced as having great potential in treating cancer metastasis [216]. In contrast to conventional therapeutic methods such as chemotherapy, surgery, and radiotherapy, the utilization of NIR light within the 700–1100 nm range to induce hyperthermia holds particular appeal. This is because biological systems generally lack chromophores that absorb within the NIR region [217].
PTT necessitates a specific agent capable of being excited and transforming electromagnetic energy into heat upon exposure to a particular light source. Ideally, an ideal PTT agent should exhibit the following properties: (1) a high rate of photothermal conversion; (2) suitable biocompatibility; and (3) straightforward conjugation capacity. GO not only fully satisfies all of these requirements but also offers additional advantageous properties that are highly beneficial in the context of cancer PTT [218].
A biocompatible platform known as porphyrin-functionalized graphene oxide (PGO) has been synthesized and designed with a strong absorption capacity at 808 nm. This PGO platform, equipped with active functional groups, enables precise targeting in PTT for brain cancer treatment while preserving the well-being of healthy cells and tissues [219]. A theranostic nanomedicine, denoted as GO-PEG(TP), has been developed, comprising PEGylated nano graphene oxide co-loaded with photosensitizers (PS) and a two-photon compound, to combat cancer. The solutions of GO, GO-PEG, and GO-PEG(TP) (each with a concentration of 0.05 mg/mL of GO) were compared to demonstrate that this integrated therapeutic approach shows remarkable efficacy in targeting and eradicating 4T1 murine breast cancer cells, resulting in a significant induction of apoptosis among the cell population [220].
Gong et al. created nanocomposites by functionalizing magnetic graphene oxide (MGO) with triformyl cholic acid and folic acid (MGO-TCA-FA). This innovative approach establishes an efficient nanoplatform for photo-chemotherapy (PCT) for targeting liver cancer. The advantages of this platform include the capability for multiple-targeted drug delivery, drug release triggered by both pH levels and NIR, and remarkable efficiency in photothermal conversion [221]. GO is combined with PEG as a photothermal material to induce a heating effect in macrophages to enable its anti-tumor effect in vitro and in vivo [222].
A photothermal therapeutic agent has been developed using reduced graphene oxide for the targeted ablation of A549 lung cancer cells. The fabrication method involves a one-step, biocompatible process utilizing Memecylon edule leaf extract for the reduction of GO, ultimately yielding polyphenol-anchored reduced graphene oxide (RGO). RGO exhibits remarkable sensitivity to NIR irradiation, allowing precise in vitro targeting of lung cancer cells and delivering cytotoxic effects [223]. The plant extract of Salvia spinosa facilitates the conversion of graphene oxide (GO) to RGO. It significantly destroyed the PC cells (Panc02-H7) after exposing RGO-loaded PC cells to laser radiation [224]. A report evidenced the effects of a combination cancer therapy including RT (doses of 2 and 4 Gy) and PTT (808 nm laser irradiation) as a radio-photothermal therapy (RPTT) for a KB oral squamous carcinoma cell line in the presence of Fe3O4@Au/reduced graphene oxide (rGO) nanostructures (NSs) at different concentrations [225].

Photodynamic Therapy

PDT, a Food and Drug Administration (FDA)-approved cancer therapy, relies on photosensitizers that generate reactive oxygen species (ROSs) when exposed to specific light, effectively killing cancer cells. This minimally invasive treatment damages tumor vasculature, triggers an immune response, and boasts specificity and repeatability. To enhance its effectiveness, nanocarriers are utilized to deliver photosensitizers directly to the tumor site, making PDT a promising and targeted cancer treatment option.
GO was conjugated to amine-terminated six-armed PEG via amide formation and then loaded with Ce6 via supramolecular π-π stacking, which showed lower singlet oxygen generation efficiency than free Ce6. GO-PEG-Ce6 significantly enhances photodynamic cancer-cell-killing efficacy by facilitating an increased cellular uptake of Ce6 through the nanographene carrier [226]. Hyaluronic acid (HA)–GO conjugates, with a high loading of photosensitizers (Ce6), the PDT efficiencies of which were remarkably improved ∼10 times more than that of free Ce6, significantly influenced the co-treatment with an excess amount of HA polymers, illustrating their active targeting to HA receptors overexpressed on cancer cells [227]. A novel hybrid of GO and hypocrellin B (HB) generated efficient singlet oxygen via irradiation to damage tumor cells [228]. The interaction of methylene blue (MB) as a photosensitizer with GO has good performance in PDT during red-light-emitting diode (LED) irradiation, which showed cell-killing potential on MDA-MB-231 breast cancer cells [229]. Folic acid (FA) and chlorin e6 (Ce6) double-functionalized GO can penetrate rapidly into cancer cells and macrophages, exhibiting good photothermal properties and a high ROS generation capacity. Moreover, a combined effect of PTT and PDT leads to a higher killing efficiency toward different types of cells involved in cancer and other diseases [230].
Guo et al. developed a drug delivery system incorporating Paclitaxel (PTX) onto PEG-modified and oxidized sodium alginate (OSA)-functionalized GO nanosheets (NSs), called PTX@GO-PEG-OSA. The photothermal conversion ability was tested via subjecting GO-PEG-OSA NSs and PTX@GO-PEG-OSA NSs (GO-PEG-OSA concentration ~ 0.1 mg/mL) to 808 nm of NIR laser irradiation. After NIR irradiation, PTX@GO-PEG-OSA could generate excessive ROS, attack mitochondrial respiratory chain complex enzymes, reduce adenosine-triphosphate (ATP) supplements for P-glycoprotein(P-gp), and effectively inhibit P-gp’s efflux pump function, thereby inducing obvious antitumor effects on gastric cancer [231]. A pluronic-based graphene oxide-methylene blue (GO-MB/PF127) nanocomposite was activated by both 808 nm NIR light and a 660 nm LED source. In this system, the GO component induced photothermal ablation of cancer cells, while methylene blue (MB) generated singlet oxygen, thereby destroying cancer cells via oxidative stress in PDT [232].
Nitrogen-doped graphene oxide dots (NGODs) can effectively produce H2O2 under white light irradiation and their H2O2 rate is proportional to the ascorbic acid (AA) concentration. This AA-supplemented PDT effectively kills lung, head and neck, colon, and oral cancer cells; however, it is highly safe for normal cells [233]. Liu et al. reported that nanoscale graphene oxide (NGO) was synthesized and then loaded with gold nanoparticles (AuNPs) and thiol polyethylene glycol folic acid (SH-PEG1000-FA). Further modifications involved incorporating the photosensitizer MB or the anticancer drug 5-fluorouracil (5-Fu). These multifunctional nanoplatforms were designed to facilitate either photothermal–photodynamic synergy (NGO-AuNPs-FA/MB) or photothermal–chemotherapy synergy (NGO-AuNPs-FA/5-Fu). When exposed to laser irradiation, they demonstrated exceptional photodynamic and photothermal properties, resulting in excellent in vitro antitumor effects [234].

3.3.3. Angiogenesis and Anti-Angiogenesis Therapy

In recent studies, GO showed angiogenesis or anti-angiogenesis properties. Mukherjee et al. demonstrated that the intracellular formation of reactive oxygen species and reactive nitrogen species as well as the activation of phospho-eNOS and phospho-Akt might be plausible mechanisms for GO and rGO-induced angiogenesis [235]. GO/polycaprolactone (PCL) nanoscaffolds are fabricated to evaluate their pro-angiogenic characteristics. The AKT-endothelial nitric oxide synthase (eNOS)-vascular endothelial growth factor (VEGF) signaling pathway might play a major role in the angiogenic process [236]. However, GO also affected the consumption of niacinamide, a precursor of energy carriers, and several amino acids involved in the regulation of angiogenesis. The combination of the physical hindrance of internalized GO aggregates, induction of oxidative stress, and alteration of some metabolic pathways leads to a significant antiangiogenic effect in primary human endothelial cells [237]. GO containing 6-gingerol (Ging) modified with chitosan (CS)-FA nanoparticles (Ging-GO-CS-FA) have pro-oxidant power against cancer cells by reducing the amount of the superoxide dismutase (SOD) gene, its average length, and the number of blood vessels on angiogenesis [238].

4. Frontiers in Pancreatic Cancer Organoids and Graphene Oxide Platform

4.1. Current Research of Graphene Oxide in 3D Culture Tumor

3D culturing represents a good model for restructuring tissue microenvironments. In recent years, more and more researchers have focused on nanoparticles and 3D culturing of spheroids or organoids. Herein, we summarize these studies, which evidence the varying effects of graphene-based nanoparticles on 3D culture cells mode (Table 2).
Graphene oxide selectively hinders the proliferation of cancer stem cells across a diverse range of tumor types. Utilizing the tumor-sphere assay, GO’s efficacy in inhibiting tumor-sphere formation was represented in six distinct cancer types, including breast, ovarian, prostate, lung, pancreatic, and glioblastoma cancers [239]. A microfluidic lab-on-a-chip system was effectively employed to establish four-day cultures of liver (HepG2), breast (MCF-7), and colon (HT-29) cancer cells in both 2D monolayer and 3D spheroid configurations. Notably, the removal of graphene oxide proved to be more straightforward from flat structures than from three-dimensional ones. The variations between 2D and 3D models may arise from distinct cellular responses contingent on their developmental context [240].
GO coated on the sidewalls of micro-wells fabricated from a cell-adhesion-resistive polymer was found to efficiently initiate the distinct donut-like formation of cancer cell spheroids. Vertically coated GO micropatterns (vGO-MPs) of varying sizes (100–250 µm) were observed on polymer platforms with HepG2 cells. The 150 µm-sized platform was found to efficiently and rapidly induce the formation of 3D spheroids in the absence of cell-adhesion proteins [241]. Grilli et al. analyzed the efficiency of 3D lung cancer spheroids combined with a minimally modified graphene oxide -based nanocarrier for siRNA delivery as a new system for cell transfection, which demonstrated the higher efficiency of spheroids compared to 2D models for transfection and the high potential of unmodified GO to carry siRNA [242].
GO flake interactions with both in vitro (U-87 MG three-dimensional spheroids, devoid of stromal or immune components) and in vivo (U-87 MG orthotopic xenograft) glioblastoma models were examined, which suggested that GO flakes can achieve a comprehensive and uniform distribution within glioblastoma tumors and serve as a potential approach to target their myeloid compartment. This opens possibilities for employing GO flakes as a platform to develop immunomodulation strategies against glioblastoma by specifically targeting macrophage and microglia compartments [243]. Graphene oxide inhibits the growth and malignancy of glioblastoma stem cell-like (GSC) spheroids through epigenetic mechanisms that drive the differentiation of GSCs, thereby reducing the malignancy of glioblastoma multiforme(GBM) [244]. GO loaded with PEG superparamagnetic iron oxide nanoparticles and grafted with methotrexate and stimuli-responsive linkers (GO-SPION-MTX) were developed for breast cancer treatment in 3D culture. These nanocomposites were internalized by cancer cells expressing folate receptors and demonstrated high cytotoxicity when subjected to NIR laser rays [245].
Table 2. The applications of graphene-based nanoparticles in 3D culture cells.
Table 2. The applications of graphene-based nanoparticles in 3D culture cells.
Type of Graphene-Based NanoparticleCategories of OriginFunction and HighlightsRef.
Graphene oxide (GO)liver (HepG2), breast (MCF-7) and colon (HT-29) cancer cellsMicrofluidic Lab-on-a-Chip systems[240]
GOhuman brainused multi-omics techniques to investigate the mechanisms of GO on lipid homeostasis in a 3D brain organoid model. Transcriptomics and lipidomics indicated that direct contact with GO altered lipid homeostasis through ER stress in 3D human brain organoids[246]
GOInner ear organoids (IEOs)promote cell–extracellular matrix interactions and cell–cell gap junctions, potential applications for drug testing[247]
GOU87, U251 GSCs and primary GSCsGO could promote differentiation and reduce malignancy in GSCs via an unanticipated epigenetic mechanism[244]
Graphene oxide (GO) flakesHuman GlioblastomaGO flakes translocated deeply into the spheroids[243]
Gold-graphene hybrid nanomaterial (Au@GO)co-culture spheroids of HeLa/Ovarian cancer and HeLa/human umbilical vein endothelial cell (HUVEC)Au@GO nanoparticles displayed selectivity towards the fast-dividing HeLa cells, which could not be observed to this extent in 2D cultures. [248]
Graphene oxide (GO)-based nanocarrier for siRNAlung cancer(A549)high potential of unmodified GO to carry siRNA[242]
Fluorescent chitosan/graphene oxide hybrid microspheres (GCS/GO)human umbilical cord mesenchymal stem cellsThe hybrid microspheres can support long-time stem cell expansion, autofluorescence also makes observing and tracking the stem cells’ behavior on the surface of microsphere scaffolds [249]
HA-EDA-PHEA-DVS/GO composite nano gel human colon cancer cells (HCT 116)conduct thermal ablation of solid tumors[250]
graphene oxide (GO) loaded with PEGylated superparamagnetic iron oxide nanoparticles and grafted with methotrexate and stimuli-responsive linkers (GO-SPION-MTX) breast cancer cell GO-SPION-MTX was internalized by the folate-receptor-positive cancer cells and induced high cytotoxicity on exposure to NIR laser rays[245]
Herceptin-stabilized graphenebreast cancer cells (SKBR-3)ultrasonic-assisted method in one-step synthesis, long-term stability in aqueous solutions[251]
Graphene nanoplateshuman endothelial cells, human brain perivascular pericytes, primary neurons, human astroglia cells, and primary microgliaspheroid bulk is formed by neural cells and microglia and the surface by endothelial cells and they upregulate key structural and functional proteins of the blood-brain barrier. These cellular constructs are utilized to preliminary screen the permeability of polymeric, metallic, and ceramic nanoparticles[252]
Graphene quantum dots (GQDs) human hepatoma cell line (HepG2 cells)The chirality of GQDs (L/D-GQDs) was modification with L/D-cysteines. L-GQDs are more effective as nanocarriers for Doxorubicin delivery[253]
Hydroxylated GQDs (OH-GQDs)mice intestinal cryptsOH-GQD treatment significantly reduced the size of the surviving intestinal organoids.[254]
reduced graphene oxide-branched polyethyleneimine-polyethylene glycol (rGO-BPEI-PEG) uniformly sized neural stem cell (NSC)-derived neurospheresPhotothermal therapy (PTT) application of brain tumor spheroids generated by the microfluidic device using rGO-BPEI-PEG nanocomposite as the PTT agent[255]
Magnetic nanoparticle-decorated reduced graphene oxide (m-rGO)neuroblastoma cells (SH-SY5Y)encapsulating SH-SY5Y to promote cell differentiation and induce oriented cell growth owing to its excellent biocompatibility and electrical conductivity[256]
Reduced graphene oxide (rGO) was the carrier for the loading of doxorubicin (DOX) and chlorin e6 (Ce6) (rGO-PEG/Ce6 and rGO-PEG/DOX)glioma cells (U87)PTT showed great treatment efficacy in the 3D tumor spheroid mode than CT and PDT[257]
Reduced graphene oxide-MXene (rGO-Mxene) hydrogelepithelial adenocarcinoma, neuroblastoma, and fibroblasts strong affinity of cellular protrusions (neurites, lamellipodia, and filopodia) to grow and connect along architectural network paths within the rGO-Mxene hydrogel, leading to control over macroscopic formations of cellular networks for technologically relevant bioengineering applications[258]
Gelatin with methacryloyl groups (GeIMA) and reduced graphene oxide (rGO) colon carcinoma cells (RKO)GelMA with higher crosslink densities and promote proliferation[259]
Vertically coated GO micropatterns (vGO-MPs)human liver cancer cells (HepG2)Cytophilic GO is selectively coated on the sidewalls of micro-wells fabricated from a cell-adhesion-resistive polymer to efficiently initiate distinct donut-like formation of cancer cell spheroids. Highly stable, the anticancer effects improved[241]
Engineered carbon nanotubes (CNTs)intestinal organoidspromoted the development of intestinal organoids over time, CNTs reduced the hardness of the extracellular matrix by decreasing the elasticity and increasing the viscosity[260]
Nanowire (NW)-templated 3D fuzzy graphene (NT-3DFG) primary E18 rat cortical tissuesa hybrid nanomaterial for remote, nongenetic, photothermal stimulation of 2D and 3D neural cellular systems.[261]
3D interconnected graphene–carbon nanotube web (3D GCNT web)glioma and healthy cortical cells3D trajectories and velocity distribution of individual infiltrating glioma to be reconstructed with unprecedented precision[262]
Three-dimensional graphene foam (3D-GFs)neural stem cell (NSC)3D-GFs can enhance the NSC differentiation towards astrocytes and especially neurons. [263]
3D-SR-Bas with active biosensors: graphene field-effect transistor (GFET)human cardiac spheroids (HUES9 hESCs)provided continuous and stable multiplexed recordings of field potentials with high sensitivity and spatiotemporal resolution,[264]

4.2. Current Research of Graphene Oxide in Pancreatic Cancer

In recent years, nanoparticles based on GO have been employed in pancreatic cancer research, demonstrating superior efficacy (Table 3). GO was employed as a gene delivery system for the simultaneous delivery of HDAC1 and K-Ras siRNAs, which target the HDAC1 gene and the G12C mutant K-Ras gene, respectively. This targeted delivery was specifically designed to affect pancreatic cancer cells, particularly MIA PaCa-2. Utilizing GO-based nanoformulations in conjunction with NIR light, tumor volume growth in mice was significantly reduced by up to 80%, which highlights the potential of GO-based nanocarriers in combining cancer gene therapy with photothermal effects for pancreatic cancer treatment [265].
Hybrid droplets containing gold–graphene oxide (Au-GO), doxorubicin, and zwitterionic chitosan (ZC) for the assembly of Au-GO@ZC-DOX stealth nanovesicles (NVs) were used to prevent macrophage opsonization, resulting in anti-cancer and anti-migration effects with high intracellular uptake in PANC-1 and MIA PaCa-2 cells [266]. The analysis of personalized biomolecular coronas (BC) with GO has demonstrated its potential for early cancer detection. In a study involving 50 subjects—half with pancreatic cancer and half being healthy volunteers—GO nanometric flakes were used to assess their BCs. The test achieved a remarkable 92% sensitivity in distinguishing cancer patients from healthy individuals, with an impressive area under the curve (AUC) of 0.96. These results highlight the promising role of GO-based BC analysis in early pancreatic cancer identification [267].
Sudhakara Prasad et al. developed affordable, disposable, paper-based immunosensors for the early quantitative detection of pancreatic cancer using the novel biomarker SGK269 (PEAK1). These immunosensors feature paper-based electrodes (PPEs) coated with GO, offering a specific diagnostic platform suitable for point-of-care and low-resource settings [268]. Graphene oxide nanosheets on a patterned gold surface were used to capture circulating tumor cells (CTCs) from blood samples of pancreatic cancer patients, as CTCs can be valuable biomarkers for disease diagnosis and progression [269].
Quagliarini et al. underscored the potential of leveraging the magnetic levitation of graphene oxide–protein complexes in conjunction with blood glucose levels for the early detection of PDAC. The observed significant variations in levitating nanosystems between controls and PDACs suggest the feasibility of employing this innovative approach as a multiplexed blood test for PDAC screening, especially in populations with hyperglycemia [270]. A positively charged lipid bilayer membrane was applied to reduced graphene oxide@gold nanostar (rGO@AuNS) for photoacoustic/photothermal dual-modal imaging-guided gene/photothermal synergistic therapy in pancreatic cancer. The combined photothermal and gene therapy, targeting the G12V mutant K-Ras gene, demonstrated remarkable anticancer and anti-liver metastasis efficacy in mice with pancreatic cancer tumors [271].
Table 3. The applications of GO-based nanoparticles in pancreatic cancer research.
Table 3. The applications of GO-based nanoparticles in pancreatic cancer research.
NanocarrierFunctionModel of
Pancreatic Cancer
Type of StudyReference
GOselectively targets cancer stem cells (CSCs) of multiple cancer cell typesMIA PaCa-2 in vitro[239]
GOa multiplexed Maglev-based nanotechnology as a screening tool for PDAC in populations with hyperglycemiaplasma of patientsin vitro[270]
GOenhance the combined effect of hyperthermia and radiation treatment-in vitro[272]
GOinvestigate toxicityBxPC-3, AsPC-1in vitro[273]
GO nanoflakescancer identification at early stages via analysis of the personalized biomolecular corona (BC)plasma of patientsin vitro[267]
GO nanosheetsnanoparticle-enabled blood test and serum levels of acute-phase protein detectionhuman blood in vitro[274]
GO sheetssynergistic analysis of protein corona and hemoglobin levelsplasma of patientsin vitro[275]
GO-Au nanosheetscirculating tumor cells (CTCs) were captured with high sensitivity at a low concentration of target cellsblood of patientin vitro[269]
GO–Protein Corona Complexesprotein cornona detection, in vitro diagnostic (IVD) testingplasma samples of PDAC patientsin vitro[276]
GOQDsfacile pulsed-laser ablation in liquid (PLAL) technique for preparing GOQDs exhibited excellent optoelectronic propertiesPANC-1in vitro[277]
GQD-HSA-Gemdrug delivery and bioimagingPANC-1in vitro[278]
MWCNT-COOH/GOuse as two- and three-dimensional scaffolds to tissue engineer tumor modelsPANC-1, BxPC-3, AsPC-1in vitro[279]
Nitrogen-doped graphene quantum dots (NGQDs)pre-miR-132 detection for diagnosis-in vitro[280]
PAH/FA/PEG/GO siRNA (HDAC1/K-Ras) complexsiRNA delivery, photothermal (808 nm), and gene therapyMIA PaCa-2/Athymic nude mice (BALB/cASlac-nu)in vitro, in vivo[265]
Polymer-GOEfficient capture and release of viable circulating tumor cellsPANC-1in vitro[281]
Reduced GO-gold-palladium (rGO-Au-Pd) detect carbohydrate antigen 24-2 (CA242) markerhuman serumin vitro[282]
rGOPhotothermal (980 nm)mice Panc02-H7/C57BL/6 mice in vitro, in vivo[283]
RGOS. spinosa leaf extract reduced the GO into RGO, photothermal (808 nm) mice Panc02-H7in vitro[224]
RGO FETidentify early diagnostic biomarker miRNA10bplasma samplesin vitro[284]
rGO@AuNS-lipid (DODAB/DOPE-FA)imaging: photoacoustic/photothermal; therapy: PTT/geneCapan-1/Capan-1 tumor-bearing nude micein vitro, in vivo[271]
ss-DNA@GoQdot@miR-141GoQdot modified and thiolated single-stranded DNA detection probes (thiol-ss-DNA) on screen-printed electrodes (SPEs) interacted with the miR-141 marker, electrochemical biosensor-in vitro[285]
Au-GO@ZC-DOX high intracellular uptake, chemo-phototherapyMIA PaCa-2, PANC-1in vitro[266]
Anti-PEAK1-GO-PPEa low-cost electrochemical immunosensor on paper for the quantitative analysis of biomarker PEAK1-in vitro[268]
AuNCs/GO GO improved the sensitivity of AuNCs-based PEC immunosensors, Glypican-1 (GPC1), antigen (CEA), and glutathione (GSH) for early diagnosisPANC-1/Balb/c-nu micein vitro, in vivo[286]
carboxyl-GOultra-sensitive carboxyl-functionalized graphene oxide (GO-COOH)-based surface plasmon resonance (SPR) immunosensors using a carbohydrate antigen (CA) 199 (CA199) biomarkerblood of the patientin vitro[287]

4.3. Limitation of Current Organoids for Pancreatic Cancer Research

The current research indicates significant progress in utilizing pancreatic cancer organoids to simulate the pathophysiological processes of pancreatic cancer. However, several limitations remain. Firstly, prolonged culturing may alter organoid behavior, potentially compromising the ability to faithfully recapitulate tissue phenotypes and pathophysiological processes because of a potential loss of genetic diversity over extended periods [288,289]. Secondly, variations in the culture media’s composition can influence organoid behavior. While organoids can maintain complex cell interactions in cultures, changes in response to different culture conditions pose challenges in accurately interpreting experimental results and achieving study reproducibility [290]. Furthermore, capturing the full heterogeneity of the disease is very complex, especially given its intricate and diverse nature. While organoids derived from pancreatic cancer patients offer valuable insights, they may not fully encompass the entire spectrum of tumor heterogeneity. This limitation can be attributed to the selective growth of certain cell types in organoids and potential changes during the culturing process [67,291,292]. Lastly, existing pancreatic cancer organoid models often lack a fully functional vascular system, limiting their ability to faithfully replicate the tumor microenvironment. The absence of a vascular system hinders a comprehensive understanding of tumor growth, invasion, and treatment responses [94,103,293].

4.4. The Barriers of Graphene Oxide for Pancreatic Cancer Research

GO also has several limitations and challenges in pancreatic cancer research that necessitate further investigation. Firstly, there is a critical need for comprehensive, long-term biosafety studies to assess the potential impact on human health. The toxicity and potential side effects of GO on healthy cells remain a concern and require thorough examination [294]. The mechanism of GO degradation and excretion is not well elucidated, and understanding these processes is essential for assessing its long-term safety, particularly as the oxide part of GO may induce the generation of reactive oxygen species (ROS), influencing cytotoxicity [295]. The sensitivity of GO’s conductivity and capacity to respond to local electrical and chemical perturbations is crucial for its effectiveness in cancer research. Although GO can absorb light in the near-infrared (NIR) region, its absorption is not sufficiently strong, raising questions about its efficacy in photothermal applications for pancreatic cancer treatment. Immunogenicity is a vital consideration [296,297]. Additionally, the water solubility of GO needs improvement, and future studies should focus on enhancing nanocarrier biocompatibility and stability, reducing size-related toxicity, and preventing agglomeration during biomedical applications [298].
GO composites, known for their durability, may persist and accumulate in environments such as water bodies and soil, potentially leading to ecological imbalances due to their limited degradability [299]. Studies indicate that these nanoparticles can be toxic to aquatic organisms, causing oxidative stress and cellular damage, and may also lead to bioaccumulation in the food chain, impacting the transport and toxicity of other environmental contaminants [300].
The preparation of GO composites faces significant challenges in achieving reproducibility and uniformity because of factors such as the degree of oxidation, defects, and variations in the size and thickness of GO sheets [301,302]. Scaling up production and integrating materials homogenously, while maintaining GO’s inherent properties and ensuring cost-effectiveness, are major hurdles for their widespread commercialization [301].
In conclusion, while GO holds promise in pancreatic cancer research, these unresolved issues and challenges underscore the importance of continued investigation and optimization for its safe and effective application in biomedical settings.

4.5. Future of Pancreatic Cancer Organoids and Graphene Oxide Model Systems

In recent years, there has been significant progress in personalized therapy for pancreatic cancer, primarily driven by advancements in organoid technology. This approach has become increasingly important, especially considering that the majority of pancreatic malignancies are pancreatic ductal adenocarcinomas (PDAC). The groundwork for this technology was laid by Clevers’ lab, which has been pivotal in the development of pancreatic tissue organoids. The pioneering work in this field was published in 2013 by Huch et al., who successfully developed pancreatic organoids from isolated pancreatic ducts in mice [303]. This breakthrough provided a foundational model for studying pancreatic diseases. Following this, in 2015, Boj et al. made a significant advancement by reporting the first tumor organoids derived from human PDAC, marking a substantial step forward in personalized cancer treatment [24].
As we discussed the current limitations of pancreatic cancer organoids in the previous section, an advanced organoids model must necessarily be improved in the future. PDO can mimic various disease phases in PDAC; however, genetic diversity could be lost over extended periods. 3D printing allows for the creation of organoids with intricate structures that closely mimic the architecture [304]. In addition, it is produced in a standardized, reproducible manner, which is important for consistent research outcomes. Scalability is also enhanced, allowing for specific scale production of pancreatic cancer organoids. The realistic structure and function of 3D-printed organoids make them ideal for more accurate drug testing and disease modeling [305,306].
GO, as nanocarriers, have delivery capabilities which enable precise dosing and the targeted delivery of therapeutics within the organoid model. This can lead to more accurate assessments of drug efficacy and toxicity for individual patients. The integration of drug screening with CRISPR-Cas9 genome editing, as explored by Hirt et al., represents a significant advance in studying drug–gene interactions in PDAC [307]. This approach enables the modification of genes associated with disease and drug resistance and the introduction of tumor-suppressing genes, showing great promise in pancreatic cancer treatment. Crucial to this method is the development of safe, non-viral vectors for efficient nucleic acid delivery, a focus of intense research interest [308]. GO has emerged as a promising candidate in this field. Its ability to load both single-stranded DNA and RNA, despite its overall negative charge, is facilitated by hydrophobic and π-π stacking interactions between nucleic acid nucleobases and GO’s hexagonal carbon lattice [309]. This interaction may be enhanced by partial deformation of the nucleic acid’s double helix, improving adsorption onto GO’s surface.
Moreover, environmental conditions such as high salt concentrations and low pH have been shown to significantly augment the binding of double-stranded nucleic acids onto GO, suggesting its potential as an effective vector in gene therapy for pancreatic cancer [310]. Natural killer (NK) cells, when combined with antibodies that induce antibody-dependent cell-mediated cytotoxicity (ADCC), present a highly promising therapeutic approach for pancreatic cancer. Beelen et al. demonstrated that the ADCC-inducing antibodies avelumab (anti-PD-L1) and trastuzumab (anti-HER2) enhance NK cell-mediated cytotoxicity, leading to increased death of organoid cells [311]. Furthermore, when combined with GO, it is possible to create a more complex and interactive model that includes not just cancer cells, but also the surrounding stroma and immune cells.
Research has investigated the photothermal effects of GO under near-infrared irradiation. GO can convert light energy into heat energy, which can be used to increase the temperature of tumor tissue, effectively killing tumor cells [312]. The extracellular matrix (ECM) plays a crucial role in this process, acting as a barrier within the tumor microenvironment. In 3D culture organoids, which mimic this structure, the penetration of GO and its ability to transport heat are essential. This heat can damage the DNA of cancer cells, making GO’s properties vital for effective photothermal therapy [313,314,315]. Accordingly, the photodynamic therapy of GO also can be explored in organoids for personalized pancreatic cancer treatment.
In addition, malignant pancreatic cancer is often subject to late diagnosis. The unique optical properties of GO, such as fluorescence quenching, can be utilized in the development of optical biosensors for pancreatic cancer detection [316]. These sensors can offer rapid and non-invasive diagnosis, which is critical for early detection of this aggressive cancer. Another study focused on the protein corona (PC) formed around GO nanoflakes in human plasma. This research has implications for early cancer detection, where changes in the concentration of typical biomarkers are often too low to be detected by blood tests. The study suggests that the personalization of PC in GO can be maximized, enhancing the profiling between cancer vs. non-cancer patients [276].
Besides these properties mentioned above, the PH sensitivity, angiogenesis, and anti-angiogenesis ability of GO also are necessary for pancreatic cancer research. For example, the microenvironment of cancer is always more acidic than other normal parts [317,318]. The vessel formation is important for the tumor to obtain nutrition. Integrating GO with PDO can potentially enhance the effectiveness of drug delivery systems, improve imaging and monitoring capabilities, and offer new therapeutic strategies such as photothermal therapy. As artificial intelligence (AI) has been developing very quickly recently, more and more complex analyses and restructure models will help to improve the function of organoids and GO. The combination of these two platforms could lead to more accurate and personalized treatments for pancreatic cancer, addressing the need for advanced and effective therapeutic approaches in PDAC.

5. Conclusions

In conclusion, this review underscored the promising potential of leveraging the synergy between organoids and GO in advancing pancreatic cancer treatment. While organoids offer a faithful replication of the cancer’s genetic diversity and microenvironment, GO provides versatile applications in targeted drug delivery and diagnostics. Despite their individual merits, both organoids and GO face limitations, such as the complexity of replicating tumor heterogeneity and the biosafety concerns of GO. Looking forward, the integration of these two platforms could revolutionize the study of pancreatic cancer, aiding in personalized medicine, effective drug screening, and biomarker discovery. However, further research is necessary to address current challenges and optimize this combined approach for clinical application, potentially offering new hope in a field marked by limited treatment options and poor prognosis.

Author Contributions

Conceptualization, S.M.; writing—original draft preparation and editing, S.M.; writing—review and editing, I.I.-S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the European Union’s Horizon 2020 research and innovation program under the Marie Sklodowska-Curie grant agreement number 861196. National Science Centre, Poland, PRELUDIUM-20 grants, Nr. 2021/41/N/NZ7/04480.

Acknowledgments

All figures have been designed with BioRender.com and Open AI (ChatGPT4 and DALL.E).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Siegel, R.L.; Miller, K.D.; Fuchs, H.E.; Jemal, A. Cancer statistics, 2021. Ca Cancer J. Clin. 2021, 71, 7–33. [Google Scholar] [CrossRef] [PubMed]
  2. World Health Organization. Who Report on Cancer: Setting Priorities, Investing Wisely and Providing Care for All; World Health Organization: Geneva, Switzerland, 2020. [Google Scholar]
  3. Luo, W.; Wang, J.; Chen, H.; Ye, L.; Qiu, J.; Liu, Y.; Wang, R.; Weng, G.; Liu, T.; Su, D.; et al. Epidemiology of pancreatic cancer: New version, new vision. Clin. J Cancer Res. 2023, 35, 438–450. [Google Scholar] [CrossRef]
  4. Kubota, K. Recent advances and limitations of surgical treatment for pancreatic cancer. World J. Clin. Oncol. 2011, 2, 225–228. [Google Scholar] [CrossRef] [PubMed]
  5. Yang, G.; Guan, W.; Cao, Z.; Guo, W.; Xiong, G.; Zhao, F.; Feng, M.; Qiu, J.; Liu, Y.; Zhang, M.Q.; et al. Integrative Genomic Analysis of Gemcitabine Resistance in Pancreatic Cancer by Patient-derived Xenograft Models. Clin. Cancer Res. 2021, 27, 3383–3396. [Google Scholar] [CrossRef] [PubMed]
  6. Zhang, B.; Zhou, F.; Hong, J.; Ng, D.M.; Yang, T.; Zhou, X.; Jin, J.; Zhou, F.; Chen, P.; Xu, Y. The role of FOLFIRINOX in metastatic pancreatic cancer: A meta-analysis. World J. Surg. Oncol. 2021, 19, 182. [Google Scholar] [CrossRef] [PubMed]
  7. Hammel, P.; Huguet, F.; van Laethem, J.-L.; Goldstein, D.; Glimelius, B.; Artru, P.; Borbath, I.; Bouché, O.; Shannon, J.; André, T.; et al. Effect of Chemoradiotherapy vs. Chemotherapy on Survival in Patients with Locally Advanced Pancreatic Cancer Controlled after 4 Months of Gemcitabine with or without Erlotinib: The LAP07 Randomized Clinical Trial. JAMA 2016, 315, 1844–1853. [Google Scholar] [CrossRef]
  8. Garcia, P.L.; Miller, A.L.; Yoon, K.J. Patient-Derived Xenograft Models of Pancreatic Cancer: Overview and Comparison with Other Types of Models. Cancers 2020, 12, 1327. [Google Scholar] [CrossRef]
  9. Huang, W.; Navarro-Serer, B.; Jeong, Y.J.; Chianchiano, P.; Xia, L.; Luchini, C.; Veronese, N.; Dowiak, C.; Ng, T.; Trujillo, M.A.; et al. Pattern of Invasion in Human Pancreatic Cancer Organoids Is Associated with Loss of SMAD4 and Clinical Outcome. Cancer Res. 2020, 80, 2804–2817. [Google Scholar] [CrossRef]
  10. Miyabayashi, K.; Baker, L.A.; Deschênes, A.; Traub, B.; Caligiuri, G.; Plenker, D.; Alagesan, B.; Belleau, P.; Li, S.; Kendall, J.; et al. Intraductal Transplantation Models of Human Pancreatic Ductal Adenocarcinoma Reveal Progressive Transition of Molecular Subtypes. Cancer Discov. 2020, 10, 1566–1589. [Google Scholar] [CrossRef]
  11. Zhang, X.; Wang, Y.; Luo, G.; Xing, M. Two-dimensional graphene family material: Assembly, biocompatibility and sensors applications. Sensors 2019, 19, 2966. [Google Scholar] [CrossRef]
  12. Aliyev, E.; Filiz, V.; Khan, M.M.; Lee, Y.J.; Abetz, C.; Abetz, V. Structural Characterization of Graphene Oxide: Surface Functional Groups and Fractionated Oxidative Debris. Nanomaterials 2019, 9, 1180. [Google Scholar] [CrossRef]
  13. Han, X.M.; Zheng, K.W.; Wang, R.L.; Yue, S.F.; Chen, J.; Zhao, Z.W.; Song, F.; Su, Y.; Ma, Q. Functionalization and optimization-strategy of graphene oxide-based nanomaterials for gene and drug delivery. Am. J. Transl. Res. 2020, 12, 1515–1534. [Google Scholar]
  14. Ma, X.; Tao, H.; Yang, K.; Feng, L.; Cheng, L.; Shi, X.; Li, Y.; Guo, L.; Liu, Z. A functionalized graphene oxide-iron oxide nanocomposite for magnetically targeted drug delivery, photothermal therapy, and magnetic resonance imaging. Nano Res. 2012, 5, 199–212. [Google Scholar] [CrossRef]
  15. Xiao, X.; Zhang, Y.; Zhou, L.; Li, B.; Gu, L. Photoluminescence and Fluorescence Quenching of Graphene Oxide: A Review. Nanomaterials 2022, 12, 2444. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, J.; Yang, Y.; Li, K.; Li, J. Application of graphene oxide in tumor targeting and tumor therapy. J. Biomater. Sci. Polym. Ed. 2023, 34, 2551–2576. [Google Scholar] [CrossRef]
  17. Pourjavadi, A.; Asgari, S.; Hosseini, S.H. Graphene oxide functionalized with oxygen-rich polymers as a pH-sensitive carrier for co-delivery of hydrophobic and hydrophilic drugs. J. Drug Deliv. Sci. Technol. 2020, 56, 101542. [Google Scholar] [CrossRef]
  18. Campbell, E.; Hasan, M.T.; Pho, C.; Callaghan, K.; Akkaraju, G.R.; Naumov, A.V. Graphene Oxide as a Multifunctional Platform for Intracellular Delivery, Imaging, and Cancer Sensing. Sci. Rep. 2019, 9, 416. [Google Scholar] [CrossRef]
  19. Lin, B.; Chen, H.; Liang, D.; Lin, W.; Qi, X.; Liu, H.; Deng, X. Acidic pH and High-H2O2 Dual Tumor Microenvironment-Responsive Nanocatalytic Graphene Oxide for Cancer Selective Therapy and Recognition. ACS Appl. Mater. Interfaces 2019, 11, 11157–11166. [Google Scholar] [CrossRef] [PubMed]
  20. Feng, L.; Li, K.; Shi, X.; Gao, M.; Liu, J.; Liu, Z. Smart pH-responsive nanocarriers based on nano-graphene oxide for combined chemo-and photothermal therapy overcoming drug resistance. Adv. Healthc. Mater. 2014, 3, 1261–1271. [Google Scholar] [CrossRef] [PubMed]
  21. Zhao, X.; Yang, L.; Li, X.; Jia, X.; Liu, L.; Zeng, J.; Guo, J.; Liu, P. Functionalized Graphene Oxide Nanoparticles for Cancer Cell-Specific Delivery of Antitumor Drug. Bioconjugate Chem. 2015, 26, 128–136. [Google Scholar] [CrossRef]
  22. Deb, A.; Andrews, N.G.; Raghavan, V. Natural polymer functionalized graphene oxide for co-delivery of anticancer drugs: In-vitro and in-vivo. Int. J. Biol. Macromol. 2018, 113, 515–525. [Google Scholar] [CrossRef]
  23. Sato, T.; Stange, D.E.; Ferrante, M.; Vries, R.G.; Van Es, J.H.; Van den Brink, S.; Van Houdt, W.J.; Pronk, A.; Van Gorp, J.; Siersema, P.D.; et al. Long-term expansion of epithelial organoids from human colon, adenoma, adenocarcinoma, and Barrett’s epithelium. Gastroenterology 2011, 141, 1762–1772. [Google Scholar] [CrossRef]
  24. Boj, S.F.; Hwang, C.I.; Baker, L.A.; Chio, I.I.C.; Engle, D.D.; Corbo, V.; Jager, M.; Ponz-Sarvise, M.; Tiriac, H.; Spector, M.S.; et al. Organoid models of human and mouse ductal pancreatic cancer. Cell 2015, 160, 324–338. [Google Scholar] [CrossRef] [PubMed]
  25. Kim, M.; Mun, H.; Sung, C.O.; Cho, E.J.; Jeon, H.-J.; Chun, S.-M.; Jung, D.J.; Shin, T.H.; Jeong, G.S.; Kim, D.K.; et al. Patient-derived lung cancer organoids as in vitro cancer models for therapeutic screening. Nat. Commun. 2019, 10, 3991. [Google Scholar] [CrossRef] [PubMed]
  26. Fujii, M.; Shimokawa, M.; Date, S.; Takano, A.; Matano, M.; Nanki, K.; Ohta, Y.; Toshimitsu, K.; Nakazato, Y.; Kawasaki, K.; et al. A Colorectal Tumor Organoid Library Demonstrates Progressive Loss of Niche Factor Requirements during Tumorigenesis. Cell Stem Cell 2016, 18, 827–838. [Google Scholar] [CrossRef]
  27. van de Wetering, M.; Francies, H.E.; Francis, J.M.; Bounova, G.; Iorio, F.; Pronk, A.; van Houdt, W.; van Gorp, J.; Taylor-Weiner, A.; Kester, L.; et al. Prospective Derivation of a Living Organoid Biobank of Colorectal Cancer Patients. Cell 2015, 161, 933–945. [Google Scholar] [CrossRef]
  28. Hill, S.J.; Decker, B.; Roberts, E.A.; Horowitz, N.S.; Muto, M.G.; Worley, M.J.; Feltmate, C.M.; Nucci, M.R.; Swisher, E.M.; Nguyen, H.; et al. Prediction of DNA Repair Inhibitor Response in Short-Term Patient-Derived Ovarian Cancer Organoids. Cancer Discov. 2018, 8, 1404–1421. [Google Scholar] [CrossRef]
  29. Kopper, O.; De Witte, C.J.; Lõhmussaar, K.; Valle-Inclan, J.E.; Hami, N.; Kester, L.; Balgobind, A.V.; Korving, J.; Proost, N.; Begthel, H.; et al. An organoid platform for ovarian cancer captures intra- and interpatient heterogeneity. Nat. Med. 2019, 25, 838–849. [Google Scholar] [CrossRef] [PubMed]
  30. Nanki, Y.; Chiyoda, T.; Hirasawa, A.; Ookubo, A.; Itoh, M.; Ueno, M.; Akahane, T.; Kameyama, K.; Yamagami, W.; Kataoka, F.; et al. Patient-derived ovarian cancer organoids capture the genomic profiles of primary tumours applicable for drug sensitivity and resistance testing. Sci. Rep. 2020, 10, 12581. [Google Scholar] [CrossRef]
  31. Drost, J.; Karthaus, W.R.; Gao, D.; Driehuis, E.; Sawyers, C.L.; Chen, Y.; Clevers, H. Organoid culture systems for prostate epithelial and cancer tissue. Nat. Protoc. 2016, 11, 347–358. [Google Scholar] [CrossRef] [PubMed]
  32. Servant, R.; Garioni, M.; Vlajnic, T.; Blind, M.; Pueschel, H.; Müller, D.C.; Zellweger, T.; Templeton, A.J.; Garofoli, A.; Maletti, S.; et al. Prostate cancer patient-derived organoids: Detailed outcome from a prospective cohort of 81 clinical specimens. J. Pathol. 2021, 254, 543–555. [Google Scholar] [CrossRef]
  33. Driehuis, E.; Gracanin, A.; Vries, R.G.J.; Clevers, H.; Boj, S.F. Establishment of Pancreatic Organoids from Normal Tissue and Tumors. STAR Protoc. 2020, 1, 100192. [Google Scholar] [CrossRef] [PubMed]
  34. Gendoo, D.M.A.; Denroche, R.E.; Zhang, A.; Radulovich, N.; Jang, G.H.; Lemire, M.; Fischer, S.; Chadwick, D.; Lungu, I.M.; Ibrahimov, E.; et al. Whole genomes define concordance of matched primary, xenograft, and organoid models of pancreas cancer. PLoS Comput. Biol. 2019, 15, e1006596. [Google Scholar] [CrossRef]
  35. Broutier, L.; Mastrogiovanni, G.; Verstegen, M.M.; Francies, H.E.; Gavarró, L.M.; Bradshaw, C.R.; Allen, G.E.; Arnes-Benito, R.; Sidorova, O.; Gaspersz, M.P.; et al. Human primary liver cancer-derived organoid cultures for disease modeling and drug screening. Nat. Med. 2017, 23, 1424–1435. [Google Scholar] [CrossRef]
  36. Li, L.; Knutsdottir, H.; Hui, K.; Weiss, M.J.; He, J.; Philosophe, B.; Cameron, A.M.; Wolfgang, C.L.; Pawlik, T.M.; Ghiaur, G.; et al. Human primary liver cancer organoids reveal intratumor and interpatient drug response heterogeneity. JCI Insight 2019, 4, e121490. [Google Scholar] [CrossRef] [PubMed]
  37. Thomas, P.B.; Perera, M.P.J.; Alinezhad, S.; Joshi, A.; Saadat, P.; Nicholls, C.; Devonport, C.P.; Calabrese, A.R.; Templeton, A.R.; Wood, J.R.; et al. Culture of Bladder Cancer Organoids as Precision Medicine Tools. J. Vis. Exp. 2021, 178, e63192. [Google Scholar] [CrossRef]
  38. Yu, L.; Li, Z.; Mei, H.; Li, W.; Chen, D.; Liu, L.; Zhang, Z.; Sun, Y.; Song, F.; Chen, W.; et al. Patient-derived organoids of bladder cancer recapitulate antigen expression profiles and serve as a personal evaluation model for CAR-T cells in vitro. Clin. Transl. Immunol. 2021, 10, e1248. [Google Scholar] [CrossRef]
  39. Choi, S.Y.; Cho, Y.-H.; Kim, D.-S.; Ji, W.; Choi, C.-M.; Lee, J.C.; Rho, J.K.; Jeong, G.S. Establishment and Long-Term Expansion of Small Cell Lung Cancer Patient-Derived Tumor Organoids. Int. J. Mol. Sci. 2021, 22, 1349. [Google Scholar] [CrossRef]
  40. Hu, Y.; Sui, X.; Song, F.; Li, Y.; Li, K.; Chen, Z.; Yang, F.; Chen, X.; Zhang, Y.; Wang, X.; et al. Lung cancer organoids analyzed on microwell arrays predict drug responses of patients within a week. Nat. Commun. 2021, 12, 2581. [Google Scholar] [CrossRef]
  41. Lee, D.; Kim, Y.; Chung, C. Scientific Validation and Clinical Application of Lung Cancer Organoids. Cells 2021, 10, 3012. [Google Scholar] [CrossRef]
  42. Nanki, K.; Toshimitsu, K.; Takano, A.; Fujii, M.; Shimokawa, M.; Ohta, Y.; Matano, M.; Seino, T.; Nishikori, S.; Ishikawa, K.; et al. Divergent Routes toward Wnt and R-spondin Niche Independency during Human Gastric Carcinogenesis. Cell 2018, 174, 856–869. [Google Scholar] [CrossRef]
  43. Seidlitz, T.; Merker, S.R.; Rothe, A.; Zakrzewski, F.; Von Neubeck, C.; Grützmann, K.; Sommer, U.; Schweitzer, C.; Schölch, S.; Uhlemann, H.; et al. Human gastric cancer modelling using organoids. Gut 2019, 68, 207–217. [Google Scholar] [CrossRef] [PubMed]
  44. Jacob, F.; Salinas, R.D.; Zhang, D.Y.; Nguyen, P.T.T.; Schnoll, J.G.; Wong, S.Z.H.; Thokala, R.; Sheikh, S.; Saxena, D.; Prokop, S.; et al. A Patient-Derived Glioblastoma Organoid Model and Biobank Recapitulates Inter- and Intra-tumoral Heterogeneity. Cell 2020, 180, 188–204. [Google Scholar] [CrossRef] [PubMed]
  45. Krieger, T.G.; Tirier, S.M.; Park, J.; Jechow, K.; Eisemann, T.; Peterziel, H.; Angel, P.; Eils, R.; Conrad, C. Modeling glioblastoma invasion using human brain organoids and single-cell transcriptomics. Neuro-Oncology 2020, 22, 1138–1149. [Google Scholar] [CrossRef]
  46. Karakasheva, T.A.; Kijima, T.; Shimonosono, M.; Maekawa, H.; Sahu, V.; Gabre, J.T.; Cruz-Acuña, R.; Giroux, V.; Sangwan, V.; Whelan, K.A.; et al. Generation and Characterization of Patient-Derived Head and Neck, Oral, and Esophageal Cancer Organoids. Curr. Protoc. Stem Cell Biol. 2020, 53, e109. [Google Scholar] [CrossRef] [PubMed]
  47. Li, X.; Francies, H.E.; Secrier, M.; Perner, J.; Miremadi, A.; Galeano-Dalmau, N.; Barendt, W.J.; Letchford, L.; Leyden, G.M.; Goffin, E.K.; et al. Organoid cultures recapitulate esophageal adenocarcinoma heterogeneity providing a model for clonality studies and precision therapeutics. Nat. Commun. 2018, 9, 2983. [Google Scholar] [CrossRef]
  48. Berg, H.F.; Hjelmeland, M.E.; Lien, H.; Espedal, H.; Fonnes, T.; Srivastava, A.; Stokowy, T.; Strand, E.; Bozickovic, O.; Stefansson, I.M.; et al. Patient-derived organoids reflect the genetic profile of endometrial tumors and predict patient prognosis. Commun. Med. 2021, 1, 20. [Google Scholar] [CrossRef]
  49. Bi, J.; Newtson, A.M.; Zhang, Y.; Devor, E.J.; Samuelson, M.I.; Thiel, K.W.; Leslie, K.K. Successful Patient-Derived Organoid Culture of Gynecologic Cancers for Disease Modeling and Drug Sensitivity Testing. Cancers 2021, 13, 2901. [Google Scholar] [CrossRef] [PubMed]
  50. Boretto, M.; Maenhoudt, N.; Luo, X.; Hennes, A.; Boeckx, B.; Bui, B.; Heremans, R.; Perneel, L.; Kobayashi, H.; Van Zundert, I.; et al. Patient-derived organoids from endometrial disease capture clinical heterogeneity and are amenable to drug screening. Nat. Cell Biol. 2019, 21, 1041–1051. [Google Scholar] [CrossRef]
  51. Collins, A.; Miles, G.J.; Wood, J.; Macfarlane, M.; Pritchard, C.; Moss, E. Patient-derived explants, xenografts and organoids: 3-dimensional patient-relevant pre-clinical models in endometrial cancer. Gynecol. Oncol. 2020, 156, 251–259. [Google Scholar] [CrossRef]
  52. Lee, S.H.; Hu, W.; Matulay, J.T.; Silva, M.V.; Owczarek, T.B.; Kim, K.; Chua, C.W.; Barlow, L.J.; Kandoth, C.; Williams, A.B.; et al. Tumor Evolution and Drug Response in Patient-Derived Organoid Models of Bladder Cancer. Cell 2018, 173, 515–528. [Google Scholar] [CrossRef] [PubMed]
  53. Gao, D.; Vela, I.; Sboner, A.; Iaquinta, P.J.; Karthaus, W.R.; Gopalan, A.; Dowling, C.; Wanjala, J.N.; Undvall, E.A.; Arora, V.K.; et al. Organoid cultures derived from patients with advanced prostate cancer. Cell 2014, 159, 176–187. [Google Scholar] [CrossRef]
  54. Maenhoudt, N.; Defraye, C.; Boretto, M.; Jan, Z.; Heremans, R.; Boeckx, B.; Hermans, F.; Arijs, I.; Cox, B.; Van Nieuwenhuysen, E.; et al. Developing Organoids from Ovarian Cancer as Experimental and Preclinical Models. Stem Cell Rep. 2020, 14, 717–729. [Google Scholar] [CrossRef]
  55. Maenhoudt, N.; Vankelecom, H. Protocol for establishing organoids from human ovarian cancer biopsies. STAR Protoc. 2021, 2, 100429. [Google Scholar] [CrossRef] [PubMed]
  56. Weeber, F.; Van De Wetering, M.; Hoogstraat, M.; Dijkstra, K.K.; Krijgsman, O.; Kuilman, T.; Gadellaa-Van Hooijdonk, C.G.M.; Van Der Velden, D.L.; Peeper, D.S.; Cuppen, E.P.J.G.; et al. Preserved genetic diversity in organoids cultured from biopsies of human colorectal cancer metastases. Proc. Natl. Acad. Sci. USA 2015, 112, 13308–13311. [Google Scholar] [CrossRef] [PubMed]
  57. Yang, C.; Xia, B.-R.; Jin, W.-L.; Lou, G. Circulating tumor cells in precision oncology: Clinical applications in liquid biopsy and 3D organoid model. Cancer Cell Int. 2019, 19, 341. [Google Scholar] [CrossRef] [PubMed]
  58. Jo, J.; Xiao, Y.; Sun, A.X.; Cukuroglu, E.; Tran, H.D.; Göke, J.; Tan, Z.Y.; Saw, T.Y.; Tan, C.P.; Lokman, H.; et al. Midbrain-like Organoids from Human Pluripotent Stem Cells Contain Functional Dopaminergic and Neuromelanin-Producing Neurons. Cell Stem Cell 2016, 19, 248–257. [Google Scholar] [CrossRef]
  59. Norrie, J.L.; Nityanandam, A.; Lai, K.; Chen, X.; Wilson, M.; Stewart, E.; Griffiths, L.; Jin, H.; Wu, G.; Orr, B.; et al. Retinoblastoma from human stem cell-derived retinal organoids. Nat. Commun. 2021, 12, 4535. [Google Scholar] [CrossRef]
  60. McCauley, H.A.; Wells, J.M. Pluripotent stem cell-derived organoids: Using principles of developmental biology to grow human tissues in a dish. Development 2017, 144, 958–962. [Google Scholar] [CrossRef]
  61. Crespo, M.; Vilar, E.; Tsai, S.-Y.; Chang, K.; Amin, S.; Srinivasan, T.; Zhang, T.; Pipalia, N.H.; Chen, H.J.; Witherspoon, M.; et al. Colonic organoids derived from human induced pluripotent stem cells for modeling colorectal cancer and drug testing. Nat. Med. 2017, 23, 878–884. [Google Scholar] [CrossRef]
  62. Lee, D.F.; Su, J.; Kim, H.S.; Chang, B.; Papatsenko, D.; Zhao, R.; Yuan, Y.; Gingold, J.; Xia, W.; Darr, H.; et al. Modeling Familial Cancer with Induced Pluripotent Stem Cells. Cell 2015, 161, 240–254. [Google Scholar] [CrossRef]
  63. Hwang, J.W.; Desterke, C.; Féraud, O.; Richard, S.; Ferlicot, S.; Verkarre, V.; Patard, J.J.; Loisel-Duwattez, J.; Foudi, A.; Griscelli, F.; et al. iPSC-Derived Embryoid Bodies as Models of c-Met-Mutated Hereditary Papillary Renal Cell Carcinoma. Int. J. Mol. Sci. 2019, 20, 4867. [Google Scholar] [CrossRef] [PubMed]
  64. Huang, L.; Desai, R.; Conrad, D.N.; Leite, N.C.; Akshinthala, D.; Lim, C.M.; Gonzalez, R.; Muthuswamy, L.B.; Gartner, Z.; Muthuswamy, S.K. Commitment and oncogene-induced plasticity of human stem cell-derived pancreatic acinar and ductal organoids. Cell Stem Cell 2021, 28, 1090–1104. [Google Scholar] [CrossRef] [PubMed]
  65. Hepburn, A.C.; Sims, C.H.C.; Buskin, A.; Heer, R. Engineering Prostate Cancer from Induced Pluripotent Stem Cells—New Opportunities to Develop Preclinical Tools in Prostate and Prostate Cancer Studies. Int. J. Mol. Sci. 2020, 21, 905. [Google Scholar] [CrossRef] [PubMed]
  66. Drost, J.; Clevers, H. Translational applications of adult stem cell-derived organoids. Development 2017, 144, 968–975. [Google Scholar] [CrossRef] [PubMed]
  67. Broutier, L.; Andersson-Rolf, A.; Hindley, C.J.; Boj, S.F.; Clevers, H.; Koo, B.K.; Huch, M. Culture and establishment of self-renewing human and mouse adult liver and pancreas 3D organoids and their genetic manipulation. Nat. Protoc. 2016, 11, 1724–1743. [Google Scholar] [CrossRef] [PubMed]
  68. Drost, J.; van Jaarsveld, R.H.; Ponsioen, B.; Zimberlin, C.; van Boxtel, R.; Buijs, A.; Sachs, N.; Overmeer, R.M.; Offerhaus, G.J.; Begthel, H.; et al. Sequential cancer mutations in cultured human intestinal stem cells. Nature 2015, 521, 43–47. [Google Scholar] [CrossRef] [PubMed]
  69. Sachs, N.; De Ligt, J.; Kopper, O.; Gogola, E.; Bounova, G.; Weeber, F.; Balgobind, A.V.; Wind, K.; Gracanin, A.; Begthel, H.; et al. A Living Biobank of Breast Cancer Organoids Captures Disease Heterogeneity. Cell 2018, 172, 373–386. [Google Scholar] [CrossRef] [PubMed]
  70. Turco, M.Y.; Gardner, L.; Hughes, J.; Cindrova-Davies, T.; Gomez, M.J.; Farrell, L.; Hollinshead, M.; Marsh, S.G.E.; Brosens, J.J.; Critchley, H.O.; et al. Long-term, hormone-responsive organoid cultures of human endometrium in a chemically defined medium. Nat. Cell Biol. 2017, 19, 568–577. [Google Scholar] [CrossRef]
  71. Gu, Q.; Tomaskovic-Crook, E.; Wallace, G.G.; Crook, J.M. 3D Bioprinting Human Induced Pluripotent Stem Cell Constructs for In Situ Cell Proliferation and Successive Multilineage Differentiation. Adv. Healthc. Mater. 2017, 6, 1700175. [Google Scholar] [CrossRef]
  72. Hiller, T.; Berg, J.; Elomaa, L.; Röhrs, V.; Ullah, I.; Schaar, K.; Dietrich, A.-C.; Al-Zeer, M.; Kurtz, A.; Hocke, A.; et al. Generation of a 3D Liver Model Comprising Human Extracellular Matrix in an Alginate/Gelatin-Based Bioink by Extrusion Bioprinting for Infection and Transduction Studies. Int. J. Mol. Sci. 2018, 19, 3129. [Google Scholar] [CrossRef]
  73. Lee, A.; Hudson, A.R.; Shiwarski, D.J.; Tashman, J.W.; Hinton, T.J.; Yerneni, S.; Bliley, J.M.; Campbell, P.G.; Feinberg, A.W. 3D bioprinting of collagen to rebuild components of the human heart. Science 2019, 365, 482–487. [Google Scholar] [CrossRef]
  74. Park, J.Y.; Shim, J.-H.; Choi, S.-A.; Jang, J.; Kim, M.; Lee, S.H.; Cho, D.-W. 3D printing technology to control BMP-2 and VEGF delivery spatially and temporally to promote large-volume bone regeneration. J. Mater. Chem. B 2015, 3, 5415–5425. [Google Scholar] [CrossRef] [PubMed]
  75. Tytgat, L.; Van Damme, L.; Ortega Arevalo, M.D.P.; Declercq, H.; Thienpont, H.; Otteveare, H.; Blondeel, P.; Dubruel, P.; Van Vlierberghe, S. Extrusion-based 3D printing of photo-crosslinkable gelatin and κ-carrageenan hydrogel blends for adipose tissue regeneration. Int. J. Biol. Macromol. 2019, 140, 929–938. [Google Scholar] [CrossRef]
  76. Yi, H.G.; Jeong, Y.H.; Kim, Y.; Choi, Y.J.; Moon, H.E.; Park, S.H.; Kang, K.S.; Bae, M.; Jang, J.; Youn, H.; et al. A bioprinted human-glioblastoma-on-a-chip for the identification of patient-specific responses to chemoradiotherapy. Nat. Biomed. Eng. 2019, 3, 509–519. [Google Scholar] [CrossRef] [PubMed]
  77. Reid, J.A.; Palmer, X.-L.; Mollica, P.A.; Northam, N.; Sachs, P.C.; Bruno, R.D. A 3D bioprinter platform for mechanistic analysis of tumoroids and chimeric mammary organoids. Sci. Rep. 2019, 9, 7466. [Google Scholar] [CrossRef]
  78. Langer, E.M.; Allen-Petersen, B.L.; King, S.M.; Kendsersky, N.D.; Turnidge, M.A.; Kuziel, G.M.; Riggers, R.; Samatham, R.; Amery, T.S.; Jacques, S.L.; et al. Modeling Tumor Phenotypes In Vitro with Three-Dimensional Bioprinting. Cell Rep. 2019, 26, 608–623. [Google Scholar] [CrossRef] [PubMed]
  79. Xu, F.; Celli, J.; Rizvi, I.; Moon, S.; Hasan, T.; Demirci, U. A three-dimensional in vitro ovarian cancer coculture model using a high-throughput cell patterning platform. Biotechnol. J. 2011, 6, 204–212. [Google Scholar] [CrossRef]
  80. Zhao, Y.; Yao, R.; Ouyang, L.; Ding, H.; Zhang, T.; Zhang, K.; Cheng, S.; Sun, W. Three-dimensional printing of Hela cells for cervical tumor model in vitro. Biofabrication 2014, 6, 035001. [Google Scholar] [CrossRef]
  81. Ma, X.; Qu, X.; Zhu, W.; Li, Y.-S.; Yuan, S.; Zhang, H.; Liu, J.; Wang, P.; Lai, C.S.E.; Zanella, F.; et al. Deterministically patterned biomimetic human iPSC-derived hepatic model via rapid 3D bioprinting. Proc. Natl. Acad. Sci. USA 2016, 113, 2206–2211. [Google Scholar] [CrossRef]
  82. Aisenbrey, E.A.; Murphy, W.L. Synthetic alternatives to Matrigel. Nat. Rev. Mater. 2020, 5, 539–551. [Google Scholar] [CrossRef] [PubMed]
  83. Kleinman, H.K.; Martin, G.R. Matrigel: Basement membrane matrix with biological activity. Semin. Cancer Biol. 2005, 15, 378–386. [Google Scholar] [CrossRef] [PubMed]
  84. Acerbi, I.; Cassereau, L.; Dean, I.; Shi, Q.; Au, A.; Park, C.; Chen, Y.Y.; Liphardt, J.; Hwang, E.S.; Weaver, V.M. Human breast cancer invasion and aggression correlates with ECM stiffening and immune cell infiltration. Integr. Biol. 2015, 7, 1120–1134. [Google Scholar] [CrossRef]
  85. Broguiere, N.; Lüchtefeld, I.; Trachsel, L.; Mazunin, D.; Rizzo, R.; Bode, J.W.; Lutolf, M.P.; Zenobi-Wong, M. Morphogenesis Guided by 3D Patterning of Growth Factors in Biological Matrices. Adv. Mater. 2020, 32, 1908299. [Google Scholar] [CrossRef]
  86. Hapach, L.A.; VanderBurgh, J.A.; Miller, J.P.; Reinhart-King, C.A. Manipulation of in vitro collagen matrix architecture for scaffolds of improved physiological relevance. Phys. Biol. 2015, 12, 061002. [Google Scholar] [CrossRef] [PubMed]
  87. Jabaji, Z.; Brinkley, G.J.; Khalil, H.A.; Sears, C.M.; Lei, N.Y.; Lewis, M.; Stelzner, M.; Martín, M.G.; Dunn, J.C.Y. Type I Collagen as an Extracellular Matrix for the In Vitro Growth of Human Small Intestinal Epithelium. PLoS ONE 2014, 9, e107814. [Google Scholar] [CrossRef] [PubMed]
  88. Xiao, W.; Zhang, R.; Sohrabi, A.; Ehsanipour, A.; Sun, S.; Liang, J.; Walthers, C.M.; Ta, L.; Nathanson, D.A.; Seidlits, S.K. Brain-Mimetic 3D Culture Platforms Allow Investigation of Cooperative Effects of Extracellular Matrix Features on Therapeutic Resistance in Glioblastoma. Cancer Res. 2018, 78, 1358–1370. [Google Scholar] [CrossRef]
  89. Gjorevski, N.; Sachs, N.; Manfrin, A.; Giger, S.; Bragina, M.E.; Ordóñez-Morán, P.; Clevers, H.; Lutolf, M.P. Designer matrices for intestinal stem cell and organoid culture. Nature 2016, 539, 560–564. [Google Scholar] [CrossRef]
  90. Capeling, M.M.; Czerwinski, M.; Huang, S.; Tsai, Y.-H.; Wu, A.; Nagy, M.S.; Juliar, B.; Sundaram, N.; Song, Y.; Han, W.M.; et al. Nonadhesive Alginate Hydrogels Support Growth of Pluripotent Stem Cell-Derived Intestinal Organoids. Stem Cell Rep. 2019, 12, 381–394. [Google Scholar] [CrossRef]
  91. Gupta, A.K.; Coburn, J.M.; Davis-Knowlton, J.; Kimmerling, E.; Kaplan, D.L.; Oxburgh, L. Scaffolding kidney organoids on silk. J. Tissue Eng. Regen. Med. 2019, 13, 812–822. [Google Scholar] [CrossRef]
  92. Broguiere, N.; Isenmann, L.; Hirt, C.; Ringel, T.; Placzek, S.; Cavalli, E.; Ringnalda, F.; Villiger, L.; Züllig, R.; Lehmann, R.; et al. Growth of Epithelial Organoids in a Defined Hydrogel. Adv. Mater. 2018, 30, 1801621. [Google Scholar] [CrossRef] [PubMed]
  93. Hunt, D.R.; Klett, K.C.; Mascharak, S.; Wang, H.; Gong, D.; Lou, J.; Li, X.; Cai, P.C.; Suhar, R.A.; Co, J.Y.; et al. Engineered Matrices Enable the Culture of Human Patient-Derived Intestinal Organoids. Adv. Sci. 2021, 8, 2004705. [Google Scholar] [CrossRef]
  94. Below, C.R.; Kelly, J.; Brown, A.; Humphries, J.D.; Hutton, C.; Xu, J.; Lee, B.Y.; Cintas, C.; Zhang, X.; Hernandez-Gordillo, V.; et al. A microenvironment-inspired synthetic three-dimensional model for pancreatic ductal adenocarcinoma organoids. Nat. Mater. 2022, 21, 110–119. [Google Scholar] [CrossRef] [PubMed]
  95. Merenda, A.; Fenderico, N.; Maurice, M.M. Wnt Signaling in 3D: Recent Advances in the Applications of Intestinal Organoids. Trends Cell Biol. 2020, 30, 60–73. [Google Scholar] [CrossRef] [PubMed]
  96. Urbischek, M.; Rannikmae, H.; Foets, T.; Ravn, K.; Hyvönen, M.; De La Roche, M. Organoid culture media formulated with growth factors of defined cellular activity. Sci. Rep. 2019, 9, 6193. [Google Scholar] [CrossRef] [PubMed]
  97. Gu, Y.; Fu, J.; Lo, P.-K.; Wang, S.; Wang, Q.; Chen, H. The effect of B27 supplement on promoting in vitro propagation of Her2/neu-transformed mammary tumorspheres. J. Biotech. Res. 2011, 3, 7. [Google Scholar]
  98. Merker, S.R.; Weitz, J.; Stange, D.E. Gastrointestinal organoids: How they gut it out. Dev. Biol. 2016, 420, 239–250. [Google Scholar] [CrossRef]
  99. Rabata, A.; Fedr, R.; Soucek, K.; Hampl, A.; Koledova, Z. 3D Cell Culture Models Demonstrate a Role for FGF and WNT Signaling in Regulation of Lung Epithelial Cell Fate and Morphogenesis. Front. Cell Dev. Biol. 2020, 8, 574. [Google Scholar] [CrossRef] [PubMed]
  100. Ma, H.-c.; Zhu, Y.-j.; Zhou, R.; Yu, Y.-y.; Xiao, Z.-z.; Zhang, H.-b. Lung cancer organoids, a promising model still with long way to go. Crit. Rev. Oncol./Hematol. 2022, 171, 103610. [Google Scholar] [CrossRef]
  101. Neal, J.T.; Li, X.; Zhu, J.; Giangarra, V.; Grzeskowiak, C.L.; Ju, J.; Liu, I.H.; Chiou, S.-H.; Salahudeen, A.A.; Smith, A.R.; et al. Organoid Modeling of the Tumor Immune Microenvironment. Cell 2018, 175, 1972–1988. [Google Scholar] [CrossRef]
  102. Zhao, H.; Jiang, E.; Shang, Z. 3D Co-culture of Cancer-Associated Fibroblast with Oral Cancer Organoids. J. Dent. Res. 2021, 100, 201–208. [Google Scholar] [CrossRef]
  103. Tsai, S.; McOlash, L.; Palen, K.; Johnson, B.; Duris, C.; Yang, Q.; Dwinell, M.B.; Hunt, B.; Evans, D.B.; Gershan, J.; et al. Development of primary human pancreatic cancer organoids, matched stromal and immune cells and 3D tumor microenvironment models. BMC Cancer 2018, 18, 335. [Google Scholar] [CrossRef]
  104. Courau, T.; Bonnereau, J.; Chicoteau, J.; Bottois, H.; Remark, R.; Assante Miranda, L.; Toubert, A.; Blery, M.; Aparicio, T.; Allez, M.; et al. Cocultures of human colorectal tumor spheroids with immune cells reveal the therapeutic potential of MICA/B and NKG2A targeting for cancer treatment. J. Immunother. Cancer 2019, 7, 74. [Google Scholar] [CrossRef] [PubMed]
  105. Dijkstra, K.K.; Cattaneo, C.M.; Weeber, F.; Chalabi, M.; Van De Haar, J.; Fanchi, L.F.; Slagter, M.; Van Der Velden, D.L.; Kaing, S.; Kelderman, S.; et al. Generation of Tumor-Reactive T Cells by Co-culture of Peripheral Blood Lymphocytes and Tumor Organoids. Cell 2018, 174, 1586–1598. [Google Scholar] [CrossRef]
  106. Truelsen, S.L.B.; Mousavi, N.; Wei, H.; Harvey, L.; Stausholm, R.; Spillum, E.; Hagel, G.; Qvortrup, K.; Thastrup, O.; Harling, H.; et al. The cancer angiogenesis co-culture assay: In vitro quantification of the angiogenic potential of tumoroids. PLoS ONE 2021, 16, e0253258. [Google Scholar] [CrossRef] [PubMed]
  107. Lim, J.T.C.; Kwang, L.G.; Ho, N.C.W.; Toh, C.C.M.; Too, N.S.H.; Hooi, L.; Benoukraf, T.; Chow, P.K.; Dan, Y.Y.; Chow, E.K.; et al. Hepatocellular carcinoma organoid co-cultures mimic angiocrine crosstalk to generate inflammatory tumor microenvironment. Biomaterials 2022, 284, 121527. [Google Scholar] [CrossRef]
  108. Bonanini, F.; Kurek, D.; Previdi, S.; Nicolas, A.; Hendriks, D.; De Ruiter, S.; Meyer, M.; Clapés Cabrer, M.; Dinkelberg, R.; García, S.B.; et al. In vitro grafting of hepatic spheroids and organoids on a microfluidic vascular bed. Angiogenesis 2022, 25, 455–470. [Google Scholar] [CrossRef] [PubMed]
  109. Kook, M.G.; Lee, S.-E.; Shin, N.; Kong, D.; Kim, D.-H.; Kim, M.-S.; Kang, H.K.; Choi, S.W.; Kang, K.-S. Generation of Cortical Brain Organoid with Vascularization by Assembling with Vascular Spheroid. Int. J. Stem Cells 2022, 15, 85–94. [Google Scholar] [CrossRef]
  110. Lim, J.; Ching, H.; Yoon, J.-K.; Jeon, N.L.; Kim, Y. Microvascularized tumor organoids-on-chips: Advancing preclinical drug screening with pathophysiological relevance. Nano Converg. 2021, 8, 12. [Google Scholar] [CrossRef]
  111. Chen, Z.; Wei, X.; Dong, S.; Han, F.; He, R.; Zhou, W. Challenges and Opportunities Associated with Platelets in Pancreatic Cancer. Front Oncol. 2022, 12, 850485. [Google Scholar] [CrossRef]
  112. Mai, S.; Inkielewicz-Stepniak, I. Pancreatic cancer and platelets crosstalk: A potential biomarker and target. Front. Cell Dev. Biol. 2021, 9, 749689. [Google Scholar] [CrossRef]
  113. Whitener, K.E.; Sheehan, P.E. Graphene synthesis. Diam. Relat. Mater. 2014, 46, 25–34. [Google Scholar] [CrossRef]
  114. Marcano, D.C.; Kosynkin, D.V.; Berlin, J.M.; Sinitskii, A.; Sun, Z.; Slesarev, A.; Alemany, L.B.; Lu, W.; Tour, J.M. Improved Synthesis of Graphene Oxide. ACS Nano 2010, 4, 4806–4814. [Google Scholar] [CrossRef] [PubMed]
  115. Zhu, Y.; Murali, S.; Cai, W.; Li, X.; Suk, J.W.; Potts, J.R.; Ruoff, R.S. Graphene and Graphene Oxide: Synthesis, Properties, and Applications. Adv. Mater. 2010, 22, 3906–3924. [Google Scholar] [CrossRef]
  116. Alam, S.N.; Sharma, N.; Kumar, L. Synthesis of Graphene Oxide (GO) by Modified Hummers Method and Its Thermal Reduction to Obtain Reduced Graphene Oxide (rGO). Graphene 2017, 6, 1–18. [Google Scholar] [CrossRef]
  117. Naik, J.P.; Sutradhar, P.; Saha, M. Molecular scale rapid synthesis of graphene quantum dots (GQDs). J. Nanostructure Chem. 2017, 7, 85–89. [Google Scholar] [CrossRef]
  118. Jiříčková, A.; Jankovský, O.; Sofer, Z.; Sedmidubský, D. Synthesis and Applications of Graphene Oxide. Materials 2022, 15, 920. [Google Scholar] [CrossRef] [PubMed]
  119. Sharma, N.; Sharma, V.; Jain, Y.; Kumari, M.; Gupta, R.; Sharma, S.K.; Sachdev, K. Synthesis and Characterization of Graphene Oxide (GO) and Reduced Graphene Oxide (rGO) for Gas Sensing Application. Macromol. Symp. 2017, 376, 1700006. [Google Scholar] [CrossRef]
  120. Sharma, H.; Mondal, S. Functionalized Graphene Oxide for Chemotherapeutic Drug Delivery and Cancer Treatment: A Promising Material in Nanomedicine. Int. J. Mol. Sci. 2020, 21, 6280. [Google Scholar] [CrossRef] [PubMed]
  121. Dreyer, D.R.; Park, S.; Bielawski, C.W.; Ruoff, R.S. The chemistry of graphene oxide. Chem. Soc. Rev. 2010, 39, 228–240. [Google Scholar] [CrossRef]
  122. Sun, L. Structure and synthesis of graphene oxide. Chin. J. Chem. Eng. 2019, 27, 2251–2260. [Google Scholar] [CrossRef]
  123. Lee, C.; Wei, X.; Kysar, J.W.; Hone, J. Measurement of the elastic properties and intrinsic strength of monolayer graphene. Science 2008, 321, 385–388. [Google Scholar] [CrossRef] [PubMed]
  124. Mouhat, F.; Coudert, F.-X.; Bocquet, M.-L. Structure and chemistry of graphene oxide in liquid water from first principles. Nat. Commun. 2020, 11, 1566. [Google Scholar] [CrossRef] [PubMed]
  125. Hu, X.; Yu, Y.; Hou, W.; Zhou, J.; Song, L. Effects of particle size and pH value on the hydrophilicity of graphene oxide. Appl. Surf. Sci. 2013, 273, 118–121. [Google Scholar] [CrossRef]
  126. Balandin, A.A.; Ghosh, S.; Bao, W.; Calizo, I.; Teweldebrhan, D.; Miao, F.; Lau, C.N. Superior thermal conductivity of single-layer graphene. Nano Lett. 2008, 8, 902–907. [Google Scholar] [CrossRef] [PubMed]
  127. Baek, S.J.; Hong, W.G.; Park, M.; Kaiser, A.B.; Kim, H.J.; Kim, B.H.; Park, Y.W. The effect of oxygen functional groups on the electrical transport behavior of a single piece multi-layered graphene oxide. Synth. Met. 2014, 191, 1–5. [Google Scholar] [CrossRef]
  128. Yang, Y.; Cao, J.; Wei, N.; Meng, D.; Wang, L.; Ren, G.; Yan, R.; Zhang, N. Thermal Conductivity of Defective Graphene Oxide: A Molecular Dynamic Study. Molecules 2019, 24, 1103. [Google Scholar] [CrossRef]
  129. Chaudhary, K.; Kumar, K.; Venkatesu, P.; Masram, D.T. Protein immobilization on graphene oxide or reduced graphene oxide surface and their applications: Influence over activity, structural and thermal stability of protein. Adv. Colloid Interface Sci. 2021, 289, 102367. [Google Scholar] [CrossRef] [PubMed]
  130. Park, O.-K.; Kim, S.-G.; You, N.-H.; Ku, B.-C.; Hui, D.; Lee, J.H. Synthesis and properties of iodo functionalized graphene oxide/polyimide nanocomposites. Compos. Part B Eng. 2014, 56, 365–371. [Google Scholar] [CrossRef]
  131. Yao, Y.; Chen, X.; Zhu, J.; Zeng, B.; Wu, Z.; Li, X. The effect of ambient humidity on the electrical properties of graphene oxide films. Nanoscale Res. Lett. 2012, 7, 363. [Google Scholar] [CrossRef]
  132. Amadei, C.A.; Stein, I.Y.; Silverberg, G.J.; Wardle, B.L.; Vecitis, C.D. Fabrication and morphology tuning of graphene oxide nanoscrolls. Nanoscale 2016, 8, 6783–6791. [Google Scholar] [CrossRef] [PubMed]
  133. Ghanbari, S.; Ahour, F.; Keshipour, S. An optical and electrochemical sensor based on l-arginine functionalized reduced graphene oxide. Sci. Rep. 2022, 12, 19398. [Google Scholar] [CrossRef]
  134. Ngqalakwezi, A.; Nkazi, D.; Seifert, G.; Ntho, T. Effects of reduction of graphene oxide on the hydrogen storage capacities of metal graphene nanocomposite. Catal. Today 2020, 358, 338–344. [Google Scholar] [CrossRef]
  135. Jia, X.; Zhang, E.; Yu, X.; Lu, B. Facile Synthesis of Copper Sulfide Nanosheet@Graphene Oxide for the Anode of Potassium-Ion Batteries. Energy Technol. 2020, 8, 1900987. [Google Scholar] [CrossRef]
  136. Pereira, G.F.L.; Fileti, E.E.; Siqueira, L.J.A. Performance of supercapacitors containing graphene oxide and ionic liquids by molecular dynamics simulations. Carbon 2023, 208, 102–110. [Google Scholar] [CrossRef]
  137. Xu, B.; Yue, S.; Sui, Z.; Zhang, X.; Hou, S.; Cao, G.; Yang, Y. What is the choice for supercapacitors: Graphene or graphene oxide? Energy Environ. Sci. 2011, 4, 2826–2830. [Google Scholar] [CrossRef]
  138. Yao, K.; Manjare, M.; Barrett, C.A.; Yang, B.; Salguero, T.T.; Zhao, Y. Nanostructured Scrolls from Graphene Oxide for Microjet Engines. J. Phys. Chem. Lett. 2012, 3, 2204–2208. [Google Scholar] [CrossRef] [PubMed]
  139. Naumov, A.V. Optical Properties of Graphene Oxide. In Graphene Oxide; John Wiley and Sons: Hoboken, NJ, USA, 2016; pp. 147–174. [Google Scholar]
  140. Wu, J.; Jia, L.; Zhang, Y.; Qu, Y.; Jia, B.; Moss, D.J. Graphene Oxide for Integrated Photonics and Flat Optics. Adv. Mater. 2021, 33, 2006415. [Google Scholar] [CrossRef]
  141. Esmaeili, Y.; Bidram, E.; Zarrabi, A.; Amini, A.; Cheng, C. Graphene oxide and its derivatives as promising In-vitro bio-imaging platforms. Sci. Rep. 2020, 10, 18052. [Google Scholar] [CrossRef]
  142. Boedtkjer, E.; Pedersen, S.F. The Acidic Tumor Microenvironment as a Driver of Cancer. Annu. Rev. Physiol. 2020, 82, 103–126. [Google Scholar] [CrossRef]
  143. Zhou, T.; Zhou, X.; Xing, D. Controlled release of doxorubicin from graphene oxide based charge-reversal nanocarrier. Biomaterials 2014, 35, 4185–4194. [Google Scholar] [CrossRef]
  144. Yang, X.; Wang, Y.; Huang, X.; Ma, Y.; Huang, Y.; Yang, R.; Duan, H.; Chen, Y. Multi-functionalized graphene oxide based anticancer drug-carrier with dual-targeting function and pH-sensitivity. J. Mater. Chem. 2011, 21, 3448–3454. [Google Scholar] [CrossRef]
  145. Borandeh, S.; Abdolmaleki, A.; Abolmaali, S.S.; Tamaddon, A.M. Synthesis, structural and in-vitro characterization of β-cyclodextrin grafted L-phenylalanine functionalized graphene oxide nanocomposite: A versatile nanocarrier for pH-sensitive doxorubicin delivery. Carbohydr. Polym. 2018, 201, 151–161. [Google Scholar] [CrossRef]
  146. Kavitha, T.; Abdi, S.I.; Park, S.Y. pH-sensitive nanocargo based on smart polymer functionalized graphene oxide for site-specific drug delivery. Phys. Chem. Chem. Phys. 2013, 15, 5176–5185. [Google Scholar] [CrossRef]
  147. Wang, P.; Huang, C.; Xing, Y.; Fang, W.; Ren, J.; Yu, H.; Wang, G. NIR-Light- and pH-Responsive Graphene Oxide Hybrid Cyclodextrin-Based Supramolecular Hydrogels. Langmuir 2019, 35, 1021–1031. [Google Scholar] [CrossRef] [PubMed]
  148. Guo, L.; Shi, H.; Wu, H.; Zhang, Y.; Wang, X.; Wu, D.; An, L.; Yang, S. Prostate cancer targeted multifunctionalized graphene oxide for magnetic resonance imaging and drug delivery. Carbon 2016, 107, 87–99. [Google Scholar] [CrossRef]
  149. Gonzalez-Rodriguez, R.; Campbell, E.; Naumov, A. Multifunctional graphene oxide/iron oxide nanoparticles for magnetic targeted drug delivery dual magnetic resonance/fluorescence imaging and cancer sensing. PLoS ONE 2019, 14, e0217072. [Google Scholar] [CrossRef]
  150. Ali, R.; Aziz, M.H.; Gao, S.; Khan, M.I.; Li, F.; Batool, T.; Shaheen, F.; Qiu, B. Graphene oxide/zinc ferrite nanocomposite loaded with doxorubicin as a potential theranostic mediu in cancer therapy and magnetic resonance imaging. Ceram. Int. 2022, 48, 10741–10750. [Google Scholar] [CrossRef]
  151. 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]
  152. Molaei, M.J. Magnetic graphene, synthesis, and applications: A review. Mater. Sci. Eng. B 2021, 272, 115325. [Google Scholar] [CrossRef]
  153. Guo, Y.; Zhang, X.; Wu, F.-G. A graphene oxide-based switch-on fluorescent probe for glutathione detection and cancer diagnosis. J. Colloid Interface Sci. 2018, 530, 511–520. [Google Scholar] [CrossRef] [PubMed]
  154. Li, R.; Gao, R.; Wang, Y.; Liu, Z.; Xu, H.; Duan, A.; Zhang, F.; Ma, L. Gastrin releasing peptide receptor targeted nano-graphene oxide for near-infrared fluorescence imaging of oral squamous cell carcinoma. Sci. Rep. 2020, 10, 11434. [Google Scholar] [CrossRef]
  155. Ma, K.; Xie, W.; Liu, W.; Wang, L.; Wang, D.; Tang, B.Z. Graphene Oxide Based Fluorescent DNA Aptasensor for Liver Cancer Diagnosis and Therapy. Adv. Funct. Mater. 2021, 31, 2102645. [Google Scholar] [CrossRef]
  156. Liu, L.; Ma, Q.; Cao, J.; Gao, Y.; Han, S.; Liang, Y.; Zhang, T.; Song, Y.; Sun, Y. Recent progress of graphene oxide-based multifunctional nanomaterials for cancer treatment. Cancer Nanotechnol. 2021, 12, 18. [Google Scholar] [CrossRef]
  157. Fu, Q.; Zhu, R.; Song, J.; Yang, H.; Chen, X. Photoacoustic Imaging: Contrast Agents and Their Biomedical Applications. Adv. Mater 2019, 31, e1805875. [Google Scholar] [CrossRef] [PubMed]
  158. Zhang, Y.; Zhang, H.; Wang, Y.; Wu, H.; Zeng, B.; Zhang, Y.; Tian, Q.; Yang, S. Hydrophilic graphene oxide/bismuth selenide nanocomposites for CT imaging, photoacoustic imaging, and photothermal therapy. J. Mater Chem. B 2017, 5, 1846–1855. [Google Scholar] [CrossRef] [PubMed]
  159. Moon, H.; Kumar, D.; Kim, H.; Sim, C.; Chang, J.-H.; Kim, J.-M.; Kim, H.; Lim, D.-K. Amplified Photoacoustic Performance and Enhanced Photothermal Stability of Reduced Graphene Oxide Coated Gold Nanorods for Sensitive Photoacoustic Imaging. ACS Nano 2015, 9, 2711–2719. [Google Scholar] [CrossRef]
  160. Wang, Z.; Sun, X.; Huang, T.; Song, J.; Wang, Y. A Sandwich Nanostructure of Gold Nanoparticle Coated Reduced Graphene Oxide for Photoacoustic Imaging-Guided Photothermal Therapy in the Second NIR Window. Front. Bioeng. Biotechnol. 2020, 8, 655. [Google Scholar] [CrossRef]
  161. Zhang, Y.; Guo, Z.; Zhu, H.; Xing, W.; Tao, P.; Shang, W.; Fu, B.; Song, C.; Hong, Y.; Dickey, M.D.; et al. Synthesis of Liquid Gallium@Reduced Graphene Oxide Core–Shell Nanoparticles with Enhanced Photoacoustic and Photothermal Performance. J. Am. Chem. Soc. 2022, 144, 6779–6790. [Google Scholar] [CrossRef]
  162. Chen, J.; Liu, C.; Zeng, G.; You, Y.; Wang, H.; Gong, X.; Zheng, R.; Kim, J.; Kim, C.; Song, L. Indocyanine Green Loaded Reduced Graphene Oxide for In Vivo Photoacoustic/Fluorescence Dual-Modality Tumor Imaging. Nanoscale Res. Lett 2016, 11, 85. [Google Scholar] [CrossRef]
  163. Okada, M.; Smith, N.I.; Palonpon, A.F.; Endo, H.; Kawata, S.; Sodeoka, M.; Fujita, K. Label-free Raman observation of cytochrome c dynamics during apoptosis. Proc. Natl. Acad. Sci. USA 2012, 109, 28–32. [Google Scholar] [CrossRef] [PubMed]
  164. Liu, Q.; Wei, L.; Wang, J.; Peng, F.; Luo, D.; Cui, R.; Niu, Y.; Qin, X.; Liu, Y.; Sun, H.; et al. Cell imaging by graphene oxide based on surface enhanced Raman scattering. Nanoscale 2012, 4, 7084–7089. [Google Scholar] [CrossRef]
  165. Zhang, Z.; Wang, M.; Gao, D.; Luo, D.; Liu, Q.; Yang, J.; Li, Y. Targeted Raman Imaging of Cells Using Graphene Oxide-Based Hybrids. Langmuir 2016, 32, 10253–10258. [Google Scholar] [CrossRef] [PubMed]
  166. Seo, S.H.; Joe, A.; Han, H.W.; Manivasagan, P.; Jang, E.S. Methylene Blue-Loaded Mesoporous Silica-Coated Gold Nanorods on Graphene Oxide for Synergistic Photothermal and Photodynamic Therapy. Pharmaceutics 2022, 14, 2242. [Google Scholar] [CrossRef]
  167. Antwi-Boasiako, A.A.; Dunn, D.; Dasary, S.S.R.; Jones, Y.K.; Barnes, S.L.; Singh, A.K. Bioconjugated graphene oxide-based Raman probe for selective identification of SKBR3 breast cancer cells. J. Raman Spectrosc. 2017, 48, 1056–1064. [Google Scholar] [CrossRef]
  168. Yang, L.; Zhen, S.J.; Li, Y.F.; Huang, C.Z. Silver nanoparticles deposited on graphene oxide for ultrasensitive surface-enhanced Raman scattering immunoassay of cancer biomarker. Nanoscale 2018, 10, 11942–11947. [Google Scholar] [CrossRef] [PubMed]
  169. Sun, B.; Wu, J.; Cui, S.; Zhu, H.; An, W.; Fu, Q.; Shao, C.; Yao, A.; Chen, B.; Shi, D. In situ synthesis of graphene oxide/gold nanorods theranostic hybrids for efficient tumor computed tomography imaging and photothermal therapy. Nano Res. 2017, 10, 37–48. [Google Scholar] [CrossRef]
  170. Li, Z.; Tian, L.; Liu, J.; Qi, W.; Wu, Q.; Wang, H.; Ali, M.C.; Wu, W.; Qiu, H. Graphene Oxide/Ag Nanoparticles Cooperated with Simvastatin as a High Sensitive X-Ray Computed Tomography Imaging Agent for Diagnosis of Renal Dysfunctions. Adv. Healthc. Mater. 2017, 6, 1700413. [Google Scholar] [CrossRef]
  171. Suslova, E.V.; Kozlov, A.P.; Shashurin, D.A.; Rozhkov, V.A.; Sotenskii, R.V.; Maximov, S.V.; Savilov, S.V.; Medvedev, O.S.; Chelkov, G.A. New Composite Contrast Agents Based on Ln and Graphene Matrix for Multi-Energy Computed Tomography. Nanomaterials 2022, 12, 4110. [Google Scholar] [CrossRef]
  172. Lalwani, G.; Sundararaj, J.L.; Schaefer, K.; Button, T.; Sitharaman, B. Synthesis, characterization, in vitro phantom imaging, and cytotoxicity of a novel graphene-based multimodal magnetic resonance imaging-X-ray computed tomography contrast agent. J. Mater. Chem. B 2014, 2, 3519–3530. [Google Scholar] [CrossRef]
  173. Dou, R.; Du, Z.; Bao, T.; Dong, X.; Zheng, X.; Yu, M.; Yin, W.; Dong, B.; Yan, L.; Gu, Z. The polyvinylpyrrolidone functionalized rGO/Bi2S3 nanocomposite as a near-infrared light-responsive nanovehicle for chemo-photothermal therapy of cancer. Nanoscale 2016, 8, 11531–11542. [Google Scholar] [CrossRef]
  174. Zhang, H.; Wu, H.; Wang, J.; Yang, Y.; Wu, D.; Zhang, Y.; Zhang, Y.; Zhou, Z.; Yang, S. Graphene oxide-BaGdF5 nanocomposites for multi-modal imaging and photothermal therapy. Biomaterials 2015, 42, 66–77. [Google Scholar] [CrossRef]
  175. Divya, K.; Jisha, M.S. Chitosan nanoparticles preparation and applications. Environ. Chem. Lett. 2018, 16, 101–112. [Google Scholar] [CrossRef]
  176. Hosseini, S.M.; Mazinani, S.; Abdouss, M.; Kalhor, H.; Kalantari, K.; Amiri, I.S.; Ramezani, Z. Designing chitosan nanoparticles embedded into graphene oxide as a drug delivery system. Polym. Bull. 2022, 79, 541–554. [Google Scholar] [CrossRef]
  177. Li, J.; Cai, C.; Li, J.; Li, J.; Li, J.; Sun, T.; Wang, L.; Wu, H.; Yu, G. Chitosan-Based Nanomaterials for Drug Delivery. Molecules 2018, 23, 2661. [Google Scholar] [CrossRef] [PubMed]
  178. Yan, T.; Zhang, H.; Huang, D.; Feng, S.; Fujita, M.; Gao, X.-D. Chitosan-Functionalized Graphene Oxide as a Potential Immunoadjuvant. Nanomaterials 2017, 7, 59. [Google Scholar] [CrossRef] [PubMed]
  179. Zalba, S.; ten Hagen, T.L.M.; Burgui, C.; Garrido, M.J. Stealth nanoparticles in oncology: Facing the PEG dilemma. J. Control. Release 2022, 351, 22–36. [Google Scholar] [CrossRef]
  180. Sun, X.; Liu, Z.; Welsher, K.; Robinson, J.T.; Goodwin, A.; Zaric, S.; Dai, H. Nano-graphene oxide for cellular imaging and drug delivery. Nano Res. 2008, 1, 203–212. [Google Scholar] [CrossRef]
  181. Zhou, T.; Zhang, B.; Wei, P.; Du, Y.; Zhou, H.; Yu, M.; Yan, L.; Zhang, W.; Nie, G.; Chen, C.; et al. Energy metabolism analysis reveals the mechanism of inhibition of breast cancer cell metastasis by PEG-modified graphene oxide nanosheets. Biomaterials 2014, 35, 9833–9843. [Google Scholar] [CrossRef]
  182. Samadian, H.; Mohammad-Rezaei, R.; Jahanban-Esfahlan, R.; Massoumi, B.; Abbasian, M.; Jafarizad, A.; Jaymand, M. A de novo theranostic nanomedicine composed of PEGylated graphene oxide and gold nanoparticles for cancer therapy. J. Mater. Res. 2020, 35, 430–441. [Google Scholar] [CrossRef]
  183. Lee, S.Y.; Kang, M.S.; Jeong, W.Y.; Han, D.-W.; Kim, K.S. Hyaluronic Acid-Based Theranostic Nanomedicines for Targeted Cancer Therapy. Cancers 2020, 12, 940. [Google Scholar] [CrossRef]
  184. Vahedi, N.; Tabandeh, F.; Mahmoudifard, M. Hyaluronic acid–graphene quantum dot nanocomposite: Potential target drug delivery and cancer cell imaging. Biotechnol. Appl. Biochem. 2022, 69, 1068–1079. [Google Scholar] [CrossRef]
  185. Basu, A.; Upadhyay, P.; Ghosh, A.; Bose, A.; Gupta, P.; Chattopadhyay, S.; Chattopadhyay, D.; Adhikary, A. Hyaluronic acid engrafted metformin loaded graphene oxide nanoparticle as CD44 targeted anti-cancer therapy for triple negative breast cancer. Biochim. Biophys. Acta BBA—Gen. Subj. 2021, 1865, 129841. [Google Scholar] [CrossRef] [PubMed]
  186. Pramanik, N.; Ranganathan, S.; Rao, S.; Suneet, K.; Jain, S.; Rangarajan, A.; Jhunjhunwala, S. A Composite of Hyaluronic Acid-Modified Graphene Oxide and Iron Oxide Nanoparticles for Targeted Drug Delivery and Magnetothermal Therapy. ACS Omega 2019, 4, 9284–9293. [Google Scholar] [CrossRef] [PubMed]
  187. Aslam, M.; Kalyar, M.A.; Raza, Z.A. Polyvinyl alcohol: A review of research status and use of polyvinyl alcohol based nanocomposites. Polym. Eng. Sci. 2018, 58, 2119–2132. [Google Scholar] [CrossRef]
  188. Muppalaneni, S.; Omidian, H. Polyvinyl Alcohol in Medicine and Pharmacy: A Perspective. J. Dev. Drugs 2013, 2, 000112. [Google Scholar] [CrossRef]
  189. Mirzaie, Z.; Reisi-Vanani, A.; Barati, M. Polyvinyl alcohol-sodium alginate blend, composited with 3D-graphene oxide as a controlled release system for curcumin. J. Drug Deliv. Sci. Technol. 2019, 50, 380–387. [Google Scholar] [CrossRef]
  190. Gholami, A.; Habibi, B.; Hosseini, S.A.; Matin, A.A.; Samadi, N. Controlled release of anticancer drugs via the magnetic magnesium iron nanoparticles modified by graphene oxide and polyvinyl alcohol: Paclitaxel and docetaxel. Nanomed. J. 2021, 8, 200–210. [Google Scholar] [CrossRef]
  191. Arkaban, H.; Barani, M.; Akbarizadeh, M.R.; Pal Singh Chauhan, N.; Jadoun, S.; Dehghani Soltani, M.; Zarrintaj, P. Polyacrylic Acid Nanoplatforms: Antimicrobial, Tissue Engineering, and Cancer Theranostic Applications. Polymers 2022, 14, 1259. [Google Scholar] [CrossRef]
  192. Torabi, M.; Yaghoobi, F.; Karimi Shervedani, R.; Kefayat, A.; Ghahremani, F.; Rashidiyan Harsini, P. Mn(II) & Gd(III) deferrioxamine complex contrast agents & temozolomide cancer prodrug immobilized on folic acid targeted graphene/polyacrylic acid nanocarrier: MRI efficiency, drug stability & interactions with cancer cells. Colloids Surf. A Physicochem. Eng. Asp. 2022, 652, 129797. [Google Scholar] [CrossRef]
  193. Sgarlata, C.; D’Urso, L.; Consiglio, G.; Grasso, G.; Satriano, C.; Forte, G. pH sensitive functionalized graphene oxide as a carrier for delivering gemcitabine: A computational approach. Comput. Theor. Chem. 2016, 1096, 1–6. [Google Scholar] [CrossRef]
  194. Bharali, D.J.; Sahoo, S.K.; Mozumdar, S.; Maitra, A. Cross-linked polyvinylpyrrolidone nanoparticles: A potential carrier for hydrophilic drugs. J. Colloid Interface Sci. 2003, 258, 415–423. [Google Scholar] [CrossRef] [PubMed]
  195. Ashjaran, M.; Babazadeh, M.; Akbarzadeh, A.; Davaran, S.; Salehi, R. Stimuli-responsive polyvinylpyrrolidone-NIPPAm-lysine graphene oxide nano-hybrid as an anticancer drug delivery on MCF7 cell line. Artif. Cells Nanomed. Biotechnol. 2019, 47, 443–454. [Google Scholar] [CrossRef]
  196. Najafabadi, A.P.; Pourmadadi, M.; Yazdian, F.; Rashedi, H.; Rahdar, A.; Díez-Pascual, A.M. pH-sensitive ameliorated quercetin delivery using graphene oxide nanocarriers coated with potential anticancer gelatin-polyvinylpyrrolidone nanoemulsion with bitter almond oil. J. Drug Deliv. Sci. Technol. 2023, 82, 104339. [Google Scholar] [CrossRef]
  197. Zhang, S.; Yang, K.; Feng, L.; Liu, Z. In vitro and in vivo behaviors of dextran functionalized graphene. Carbon 2011, 49, 4040–4049. [Google Scholar] [CrossRef]
  198. Xie, M.; Lei, H.; Zhang, Y.; Xu, Y.; Shen, S.; Ge, Y.; Li, H.; Xie, J. Non-covalent modification of graphene oxide nanocomposites with chitosan/dextran and its application in drug delivery. RSC Adv. 2016, 6, 9328–9337. [Google Scholar] [CrossRef]
  199. Rajeev, M.R.; Manjusha, V.; Anirudhan, T.S. Transdermal delivery of doxorubicin and methotrexate from polyelectrolyte three layer nanoparticle of graphene oxide/polyethyleneimine/dextran sulphate for chemotherapy: In vitro and in vivo studies. Chem. Eng. J. 2023, 466, 143244. [Google Scholar] [CrossRef]
  200. Ou, L.; Sun, T.; Liu, M.; Zhang, Y.; Zhou, Z.; Zhan, X.; Lu, L.; Zhao, Q.; Lai, R.; Shao, L. Efficient miRNA Inhibitor Delivery with Graphene Oxide-Polyethylenimine to Inhibit Oral Squamous Cell Carcinoma. Int. J. Nanomed. 2020, 15, 1569–1583. [Google Scholar] [CrossRef]
  201. Hoseini-Ghahfarokhi, M.; Mirkiani, S.; Mozaffari, N.; Abdolahi Sadatlu, M.A.; Ghasemi, A.; Abbaspour, S.; Akbarian, M.; Farjadain, F.; Karimi, M. Applications of Graphene and Graphene Oxide in Smart Drug/Gene Delivery: Is the World Still Flat? Int. J. Nanomed. 2020, 15, 9469–9496. [Google Scholar] [CrossRef]
  202. Du, S.; Wang, Y.; Ao, J.; Wang, K.; Zhang, Z.; Yang, L.; Liang, X. Targeted Delivery of siRNA to Ovarian Cancer Cells Using Functionalized Graphene Oxide. Nano LIFE 2018, 8, 1850001. [Google Scholar] [CrossRef]
  203. Wang, Y.; Sun, G.; Gong, Y.; Zhang, Y.; Liang, X.; Yang, L. Functionalized Folate-Modified Graphene Oxide/PEI siRNA Nanocomplexes for Targeted Ovarian Cancer Gene Therapy. Nanoscale Res. Lett. 2020, 15, 57. [Google Scholar] [CrossRef] [PubMed]
  204. Cao, X.; Zheng, S.; Zhang, S.; Wang, Y.; Yang, X.; Duan, H.; Huang, Y.; Chen, Y. Functionalized Graphene Oxide with Hepatocyte Targeting as Anti-Tumor Drug and Gene Intracellular Transporters. J. Nanosci. Nanotechnol. 2015, 15, 2052–2059. [Google Scholar] [CrossRef] [PubMed]
  205. Qu, Y.; Sun, F.; He, F.; Yu, C.; Lv, J.; Zhang, Q.; Liang, D.; Yu, C.; Wang, J.; Zhang, X.; et al. Glycyrrhetinic acid-modified graphene oxide mediated siRNA delivery for enhanced liver-cancer targeting therapy. Eur. J. Pharm. Sci. 2019, 139, 105036. [Google Scholar] [CrossRef] [PubMed]
  206. Sun, Q.; Wang, X.; Cui, C.; Li, J.; Wang, Y. Doxorubicin and anti-VEGF siRNA co-delivery via nano-graphene oxide for enhanced cancer therapy in vitro and in vivo. Int. J. Nanomed. 2018, 13, 3713–3728. [Google Scholar] [CrossRef] [PubMed]
  207. Mirzaie, V.; Ansari, M.; Nematollahi-Mahani, S.N.; Moballegh Nasery, M.; Karimi, B.; Eslaminejad, T.; Pourshojaei, Y. Nano-Graphene Oxide-supported APTES-Spermine, as Gene Delivery System, for Transfection of pEGFP-p53 into Breast Cancer Cell Lines. Drug Des. Dev. Ther. 2020, 14, 3087–3097. [Google Scholar] [CrossRef] [PubMed]
  208. Di Santo, R.; Quagliarini, E.; Palchetti, S.; Pozzi, D.; Palmieri, V.; Perini, G.; Papi, M.; Capriotti, A.L.; Laganà, A.; Caracciolo, G. Microfluidic-generated lipid-graphene oxide nanoparticles for gene delivery. Appl. Phys. Lett. 2019, 114, 2733–2741. [Google Scholar] [CrossRef]
  209. Bugárová, N.; Špitálsky, Z.; Mičušík, M.; Bodík, M.; Šiffalovič, P.; Koneracká, M.; Závišová, V.; Kubovčíková, M.; Kajanová, I.; Zaťovičová, M.; et al. A Multifunctional Graphene Oxide Platform for Targeting Cancer. Cancers 2019, 11, 753. [Google Scholar] [CrossRef]
  210. Tran, A.V.; Shim, K.; Vo Thi, T.T.; Kook, J.K.; An, S.S.A.; Lee, S.W. Targeted and controlled drug delivery by multifunctional mesoporous silica nanoparticles with internal fluorescent conjugates and external polydopamine and graphene oxide layers. Acta Biomater. 2018, 74, 397–413. [Google Scholar] [CrossRef]
  211. Chen, S.; Zhang, S.; Wang, Y.; Yang, X.; Yang, H.; Cui, C. Anti-EpCAM functionalized graphene oxide vector for tumor targeted siRNA delivery and cancer therapy. Asian J. Pharm. Sci. 2021, 16, 598–611. [Google Scholar] [CrossRef]
  212. Li, Y.; Lu, Q.; Liu, H.; Wang, J.; Zhang, P.; Liang, H.; Jiang, L.; Wang, S. Antibody-Modified Reduced Graphene Oxide Films with Extreme Sensitivity to Circulating Tumor Cells. Adv. Mater. 2015, 27, 6848–6854. [Google Scholar] [CrossRef]
  213. Xiao, H.; Jensen, P.E.; Chen, X. Elimination of Osteosarcoma by Necroptosis with Graphene Oxide-Associated Anti-HER2 Antibodies. Int. J. Mol. Sci. 2019, 20, 4360. [Google Scholar] [CrossRef]
  214. Sahne, F.; Mohammadi, M.; Najafpour, G.D. Single-Layer Assembly of Multifunctional Carboxymethylcellulose on Graphene Oxide Nanoparticles for Improving in Vivo Curcumin Delivery into Tumor Cells. ACS Biomater. Sci. Eng. 2019, 5, 2595–2609. [Google Scholar] [CrossRef]
  215. Yang, D.; Feng, L.; Dougherty, C.A.; Luker, K.E.; Chen, D.; Cauble, M.A.; Banaszak Holl, M.M.; Luker, G.D.; Ross, B.D.; Liu, Z.; et al. In vivo targeting of metastatic breast cancer via tumor vasculature-specific nano-graphene oxide. Biomaterials 2016, 104, 361–371. [Google Scholar] [CrossRef] [PubMed]
  216. Zou, L.; Wang, H.; He, B.; Zeng, L.; Tan, T.; Cao, H.; He, X.; Zhang, Z.; Guo, S.; Li, Y. Current Approaches of Photothermal Therapy in Treating Cancer Metastasis with Nanotherapeutics. Theranostics 2016, 6, 762–772. [Google Scholar] [CrossRef] [PubMed]
  217. Feng, L.; Wu, L.; Qu, X. New horizons for diagnostics and therapeutic applications of graphene and graphene oxide. Adv. Mater. 2013, 25, 168–186. [Google Scholar] [CrossRef] [PubMed]
  218. Alemi, F.; Zarezadeh, R.; Sadigh, A.R.; Hamishehkar, H.; Rahimi, M.; Majidinia, M.; Asemi, Z.; Ebrahimi-Kalan, A.; Yousefi, B.; Rashtchizadeh, N. Graphene oxide and reduced graphene oxide: Efficient cargo platforms for cancer theranostics. J. Drug Deliv. Sci. Technol. 2020, 60, 101974. [Google Scholar] [CrossRef]
  219. Su, S.; Wang, J.; Wei, J.; Martínez-Zaguilán, R.; Qiu, J.; Wang, S. Efficient photothermal therapy of brain cancer through porphyrin functionalized graphene oxide. New J. Chem. 2015, 39, 5743–5749. [Google Scholar] [CrossRef]
  220. Liu, J.; Yuan, X.; Deng, L.; Yin, Z.; Tian, X.; Bhattacharyya, S.; Liu, H.; Luo, Y.; Luo, L. Graphene oxide activated by 980 nm laser for cascading two-photon photodynamic therapy and photothermal therapy against breast cancer. Appl. Mater. Today 2020, 20, 100665. [Google Scholar] [CrossRef]
  221. Gong, T.; Wang, X.; Ma, Q.; Li, J.; Li, M.; Huang, Y.; Liang, W.; Su, D.; Guo, R. Triformyl cholic acid and folic acid functionalized magnetic graphene oxide nanocomposites: Multiple-targeted dual-modal synergistic chemotherapy/photothermal therapy for liver cancer. J. Inorg. Biochem. 2021, 223, 111558. [Google Scholar] [CrossRef]
  222. Deng, X.; Liang, H.; Yang, W.; Shao, Z. Polarization and function of tumor-associated macrophages mediate graphene oxide-induced photothermal cancer therapy. J. Photochem. Photobiol. B Biol. 2020, 208, 111913. [Google Scholar] [CrossRef]
  223. Wang, C.; Wang, X.; Chen, Y.; Fang, Z. In-vitro photothermal therapy using plant extract polyphenols functionalized graphene sheets for treatment of lung cancer. J. Photochem. Photobiol. B Biol. 2020, 204, 111587. [Google Scholar] [CrossRef]
  224. Yang, J.; Xia, X.; He, K.; Zhang, M.; Qin, S.; Luo, M.; Wu, L. Green synthesis of reduced graphene oxide (RGO) using the plant extract of Salvia spinosa and evaluation of photothermal effect on pancreatic cancer cells. J. Mol. Struct. 2021, 1245, 131064. [Google Scholar] [CrossRef]
  225. Shakerian Ardakani, T.; Meidanchi, A.; Shokri, A.; Shakeri-Zadeh, A. Fe3O4@Au/reduced graphene oxide nanostructures: Combinatorial effects of radiotherapy and photothermal therapy on oral squamous carcinoma KB cell line. Ceram. Int. 2020, 46, 28676–28685. [Google Scholar] [CrossRef]
  226. Tian, B.; Wang, C.; Zhang, S.; Feng, L.; Liu, Z. Photothermally Enhanced Photodynamic Therapy Delivered by Nano-Graphene Oxide. ACS Nano 2011, 5, 7000–7009. [Google Scholar] [CrossRef] [PubMed]
  227. Li, F.; Park, S.-J.; Ling, D.; Park, W.; Han, J.Y.; Na, K.; Char, K. Hyaluronic acid-conjugated graphene oxide/photosensitizer nanohybrids for cancer targeted photodynamic therapy. J. Mater. Chem. B 2013, 1, 1678–1686. [Google Scholar] [CrossRef]
  228. Zhou, L.; Jiang, H.; Wei, S.; Ge, X.; Zhou, J.; Shen, J. High-efficiency loading of hypocrellin B on graphene oxide for photodynamic therapy. Carbon 2012, 50, 5594–5604. [Google Scholar] [CrossRef]
  229. Hosseinzadeh, R.; Khorsandi, K.; Hosseinzadeh, G. Graphene oxide-methylene blue nanocomposite in photodynamic therapy of human breast cancer. J. Biomol. Struct. Dyn. 2018, 36, 2216–2223. [Google Scholar] [CrossRef]
  230. Guo, S.; Song, Z.; Ji, D.-K.; Reina, G.; Fauny, J.-D.; Nishina, Y.; Ménard-Moyon, C.; Bianco, A. Combined Photothermal and Photodynamic Therapy for Cancer Treatment Using a Multifunctional Graphene Oxide. Pharmaceutics 2022, 14, 1365. [Google Scholar] [CrossRef]
  231. Guo, W.; Chen, Z.; Feng, X.; Shen, G.; Huang, H.; Liang, Y.; Zhao, B.; Li, G.; Hu, Y. Graphene oxide (GO)-based nanosheets with combined chemo/photothermal/photodynamic therapy to overcome gastric cancer (GC) paclitaxel resistance by reducing mitochondria-derived adenosine-triphosphate (ATP). J. Nanobiotechnol. 2021, 19, 146. [Google Scholar] [CrossRef]
  232. Ma, M.; Cheng, L.; Zhao, A.; Zhang, H.; Zhang, A. Pluronic-based graphene oxide-methylene blue nanocomposite for photodynamic/photothermal combined therapy of cancer cells. Photodiagnosis Photodyn. Ther. 2020, 29, 101640. [Google Scholar] [CrossRef]
  233. Shih, C.-Y.; Huang, W.-L.; Chiang, I.T.; Su, W.-C.; Teng, H. Biocompatible hole scavenger–assisted graphene oxide dots for photodynamic cancer therapy. Nanoscale 2021, 13, 8431–8441. [Google Scholar] [CrossRef]
  234. Liu, F.; Xu, C.; Li, J.; Zhang, Z.; Jin, X.; Wang, B. Nanoarchitectonics of versatile platform based on graphene oxide for precise and enhanced synergistic cancer photothermal-photodynamic/chemotherapy. J. Mol. Struct. 2023, 1294, 136499. [Google Scholar] [CrossRef]
  235. Mukherjee, S.; Sriram, P.; Barui, A.K.; Nethi, S.K.; Veeriah, V.; Chatterjee, S.; Suresh, K.I.; Patra, C.R. Graphene Oxides Show Angiogenic Properties. Adv. Healthc. Mater. 2015, 4, 1722–1732. [Google Scholar] [CrossRef]
  236. Qian, Y.; Song, J.; Zhao, X.; Chen, W.; Ouyang, Y.; Yuan, W.; Fan, C. 3D Fabrication with Integration Molding of a Graphene Oxide/Polycaprolactone Nanoscaffold for Neurite Regeneration and Angiogenesis. Adv. Sci. 2018, 5, 1700499. [Google Scholar] [CrossRef]
  237. Cibecchini, G.; Veronesi, M.; Catelani, T.; Bandiera, T.; Guarnieri, D.; Pompa, P.P. Antiangiogenic Effect of Graphene Oxide in Primary Human Endothelial Cells. ACS Appl. Mater. Interfaces 2020, 12, 22507–22518. [Google Scholar] [CrossRef] [PubMed]
  238. Al-Janabi, A.H.A.; Hayati Roodbari, N.; Homayouni Tabrizi, M. Investigating the anticancer and anti-angiogenic effects of graphene oxide nanoparticles containing 6-gingerol modified with chitosan and folate. Cancer Nanotechnol. 2023, 14, 69. [Google Scholar] [CrossRef]
  239. Fiorillo, M.; Verre, A.F.; Iliut, M.; Peiris-Pagés, M.; Ozsvari, B.; Gandara, R.; Cappello, A.R.; Sotgia, F.; Vijayaraghavan, A.; Lisanti, M.P. Graphene oxide selectively targets cancer stem cells, across multiple tumor types: Implications for non-toxic cancer treatment, via “differentiation-based nano-therapy”. Oncotarget 2015, 6, 3553–3562. [Google Scholar] [CrossRef]
  240. Zuchowska, A.; Buta, A.; Dabrowski, B.; Jastrzebska, E.; Zukowski, K.; Brzozka, Z. 3D and 2D cell models in a novel microfluidic tool for evaluation of highly chemically and microbiologically pure graphene oxide (GO) as an effective drug carrier. Sens. Actuators B Chem. 2020, 302, 127064. [Google Scholar] [CrossRef]
  241. Kim, C.-H.; Suhito, I.R.; Angeline, N.; Han, Y.; Son, H.; Luo, Z.; Kim, T.-H. Vertically Coated Graphene Oxide Micro-Well Arrays for Highly Efficient Cancer Spheroid Formation and Drug Screening. Adv. Healthc. Mater. 2020, 9, 1901751. [Google Scholar] [CrossRef] [PubMed]
  242. Grilli, F.; Hassan, E.M.; Variola, F.; Zou, S. Harnessing graphene oxide nanocarriers for siRNA delivery in a 3D spheroid model of lung cancer. Biomater. Sci. 2023, 11, 6635–6649. [Google Scholar] [CrossRef]
  243. De Lázaro, I.; Sharp, P.; Gurcan, C.; Ceylan, A.; Stylianou, M.; Kisby, T.; Chen, Y.; Vranic, S.; Barr, K.; Taheri, H.; et al. Deep Tissue Translocation of Graphene Oxide Sheets in Human Glioblastoma 3D Spheroids and an Orthotopic Xenograft Model. Adv. Ther. 2021, 4, 2000109. [Google Scholar] [CrossRef]
  244. Wang, X.; Zhou, W.; Li, X.; Ren, J.; Ji, G.; Du, J.; Tian, W.; Liu, Q.; Hao, A. Graphene oxide suppresses the growth and malignancy of glioblastoma stem cell-like spheroids via epigenetic mechanisms. J. Transl. Med. 2020, 18, 200. [Google Scholar] [CrossRef] [PubMed]
  245. Dolatkhah, M.; Hashemzadeh, N.; Barar, J.; Adibkia, K.; Aghanejad, A.; Barzegar-Jalali, M.; Omidian, H.; Omidi, Y. Stimuli-responsive graphene oxide and methotrexate-loaded magnetic nanoparticles for breast cancer-targeted therapy. Nanomedicine 2021, 16, 2155–2174. [Google Scholar] [CrossRef]
  246. Liu, X.; Yang, C.; Chen, P.; Zhang, L.; Cao, Y. The uses of transcriptomics and lipidomics indicated that direct contact with graphene oxide altered lipid homeostasis through ER stress in 3D human brain organoids. Sci. Total Environ. 2022, 849, 157815. [Google Scholar] [CrossRef] [PubMed]
  247. Park, S.; Kim, Y.J.; Sharma, H.; Kim, D.; Gwon, Y.; Kim, W.; Park, S.; Ha, C.W.; Choung, Y.H.; Kim, J. Graphene Hybrid Inner Ear Organoid with Enhanced Maturity. Nano Lett. 2023, 23, 5573–5580. [Google Scholar] [CrossRef]
  248. Lee, J.M.; Park, D.Y.; Yang, L.; Kim, E.-J.; Ahrberg, C.D.; Lee, K.-B.; Chung, B.G. Generation of uniform-sized multicellular tumor spheroids using hydrogel microwells for advanced drug screening. Sci. Rep. 2018, 8, 17145. [Google Scholar] [CrossRef] [PubMed]
  249. Zhang, S.; Ma, B.; Wang, S.; Duan, J.; Qiu, J.; Li, D.; Sang, Y.; Ge, S.; Liu, H. Mass-production of fluorescent chitosan/graphene oxide hybrid microspheres for in vitro 3D expansion of human umbilical cord mesenchymal stem cells. Chem. Eng. J. 2018, 331, 675–684. [Google Scholar] [CrossRef]
  250. Fiorica, C.; Mauro, N.; Pitarresi, G.; Scialabba, C.; Palumbo, F.S.; Giammona, G. Double-Network-Structured Graphene Oxide-Containing Nanogels as Photothermal Agents for the Treatment of Colorectal Cancer. Biomacromolecules 2017, 18, 1010–1018. [Google Scholar] [CrossRef]
  251. Askari, E.; Naghib, S.M.; Seyfoori, A.; Maleki, A.; Rahmanian, M. Ultrasonic-assisted synthesis and in vitro biological assessments of a novel herceptin-stabilized graphene using three dimensional cell spheroid. Ultrason. Sonochem. 2019, 58, 104615. [Google Scholar] [CrossRef]
  252. Kumarasamy, M.; Sosnik, A. Heterocellular spheroids of the neurovascular blood-brain barrier as a platform for personalized nanoneuromedicine. iScience 2021, 24, 102183. [Google Scholar] [CrossRef]
  253. Jeon, H.; Zhu, R.; Kim, G.; Wang, Y. Chirality-enhanced transport and drug delivery of graphene nanocarriers to tumor-like cellular spheroid. Front Chem. 2023, 11, 1207579. [Google Scholar] [CrossRef]
  254. Yu, L.; Tian, X.; Gao, D.; Lang, Y.; Zhang, X.-X.; Yang, C.; Gu, M.-M.; Shi, J.; Zhou, P.-K.; Shang, Z.-F. Oral administration of hydroxylated-graphene quantum dots induces intestinal injury accompanying the loss of intestinal stem cells and proliferative progenitor cells. Nanotoxicology 2019, 13, 1409–1421. [Google Scholar] [CrossRef]
  255. Lee, J.M.; Choi, J.W.; Ahrberg, C.D.; Choi, H.W.; Ha, J.H.; Mun, S.G.; Mo, S.J.; Chung, B.G. Generation of tumor spheroids using a droplet-based microfluidic device for photothermal therapy. Microsyst. Nanoeng. 2020, 6, 52. [Google Scholar] [CrossRef] [PubMed]
  256. Santhosh, M.; Choi, J.-H.; Choi, J.-W. Magnetic-Assisted Cell Alignment within a Magnetic Nanoparticle-Decorated Reduced Graphene Oxide/Collagen 3D Nanocomposite Hydrogel. Nanomaterials 2019, 9, 1293. [Google Scholar] [CrossRef] [PubMed]
  257. Liu, J.; Liu, K.; Feng, L.; Liu, Z.; Xu, L. Comparison of nanomedicine-based chemotherapy, photodynamic therapy and photothermal therapy using reduced graphene oxide for the model system. Biomater. Sci. 2017, 5, 331–340. [Google Scholar] [CrossRef] [PubMed]
  258. Wychowaniec, J.K.; Litowczenko, J.; Tadyszak, K.; Natu, V.; Aparicio, C.; Peplińska, B.; Barsoum, M.W.; Otyepka, M.; Scheibe, B. Unique cellular network formation guided by heterostructures based on reduced graphene oxide—Ti(3)C(2)T(x) MXene hydrogels. Acta Biomater. 2020, 115, 104–115. [Google Scholar] [CrossRef] [PubMed]
  259. Souza, S.O.L.D.; Oliveira, S.M.D.; Lehman, C.P.; Silva, M.C.D.; Silva, L.M.; Oréfice, R.L. Tuning the structure and properties of cell-embedded gelatin hydrogels for tumor organoids. Polímeros 2023, 33, e20230014. [Google Scholar] [CrossRef]
  260. Bao, L.; Cui, X.; Wang, X.; Wu, J.; Guo, M.; Yan, N.; Chen, C. Carbon Nanotubes Promote the Development of Intestinal Organoids through Regulating Extracellular Matrix Viscoelasticity and Intracellular Energy Metabolism. ACS Nano 2021, 15, 15858–15873. [Google Scholar] [CrossRef]
  261. Rastogi, S.K.; Garg, R.; Scopelliti, M.G.; Pinto, B.I.; Hartung, J.E.; Kim, S.; Murphey, C.G.E.; Johnson, N.; San Roman, D.; Bezanilla, F.; et al. Remote nongenetic optical modulation of neuronal activity using fuzzy graphene. Proc. Natl. Acad. Sci. USA 2020, 117, 13339–13349. [Google Scholar] [CrossRef]
  262. Xiao, M.; Li, X.; Song, Q.; Zhang, Q.; Lazzarino, M.; Cheng, G.; Ulloa Severino, F.P.; Torre, V. A Fully 3D Interconnected Graphene-Carbon Nanotube Web Allows the Study of Glioma Infiltration in Bioengineered 3D Cortex-Like Networks. Adv. Mater. 2018, 30, e1806132. [Google Scholar] [CrossRef]
  263. Li, N.; Zhang, Q.; Gao, S.; Song, Q.; Huang, R.; Wang, L.; Liu, L.; Dai, J.; Tang, M.; Cheng, G. Three-dimensional graphene foam as a biocompatible and conductive scaffold for neural stem cells. Sci. Rep. 2013, 3, 1604. [Google Scholar] [CrossRef]
  264. Kalmykov, A.; Huang, C.; Bliley, J.; Shiwarski, D.; Tashman, J.; Abdullah, A.; Rastogi, S.K.; Shukla, S.; Mataev, E.; Feinberg, A.W.; et al. Organ-on-e-chip: Three-dimensional self-rolled biosensor array for electrical interrogations of human electrogenic spheroids. Sci. Adv. 2019, 5, eaax0729. [Google Scholar] [CrossRef]
  265. Yin, F.; Hu, K.; Chen, Y.; Yu, M.; Wang, D.; Wang, Q.; Yong, K.-T.; Lu, F.; Liang, Y.; Li, Z. SiRNA Delivery with PEGylated Graphene Oxide Nanosheets for Combined Photothermal and Genetherapy for Pancreatic Cancer. Theranostics 2017, 7, 1133–1148. [Google Scholar] [CrossRef] [PubMed]
  266. Thapa, R.K.; Ku, S.K.; Choi, H.-G.; Yong, C.S.; Byeon, J.H.; Kim, J.O. Vibrating droplet generation to assemble zwitterion-coated gold-graphene oxide stealth nanovesicles for effective pancreatic cancer chemo-phototherapy. Nanoscale 2018, 10, 1742–1749. [Google Scholar] [CrossRef] [PubMed]
  267. Papi, M.; Palmieri, V.; Digiacomo, L.; Giulimondi, F.; Palchetti, S.; Ciasca, G.; Perini, G.; Caputo, D.; Cartillone, M.C.; Cascone, C.; et al. Converting the personalized biomolecular corona of graphene oxide nanoflakes into a high-throughput diagnostic test for early cancer detection. Nanoscale 2019, 11, 15339–15346. [Google Scholar] [CrossRef] [PubMed]
  268. Prasad, K.S.; Cao, X.; Gao, N.; Jin, Q.; Sanjay, S.T.; Henao-Pabon, G.; Li, X. A low-cost nanomaterial-based electrochemical immunosensor on paper for high-sensitivity early detection of pancreatic cancer. Sens. Actuators B Chem. 2020, 305, 127516. [Google Scholar] [CrossRef]
  269. Yoon, H.J.; Kim, T.H.; Zhang, Z.; Azizi, E.; Pham, T.M.; Paoletti, C.; Lin, J.; Ramnath, N.; Wicha, M.S.; Hayes, D.F.; et al. Sensitive capture of circulating tumour cells by functionalized graphene oxide nanosheets. Nat. Nanotechnol. 2013, 8, 735–741. [Google Scholar] [CrossRef]
  270. Quagliarini, E.; Caputo, D.; Cammarata, R.; Caracciolo, G.; Pozzi, D. Coupling magnetic levitation of graphene oxide–protein complexes with blood levels of glucose for early detection of pancreatic adenocarcinoma. Cancer Nanotechnol. 2023, 14, 16. [Google Scholar] [CrossRef]
  271. Jia, X.; Xu, W.; Ye, Z.; Wang, Y.; Dong, Q.; Wang, E.; Li, D.; Wang, J. Functionalized Graphene@Gold Nanostar/Lipid for Pancreatic Cancer Gene and Photothermal Synergistic Therapy under Photoacoustic/Photothermal Imaging Dual-Modal Guidance. Small 2020, 16, 2003707. [Google Scholar] [CrossRef]
  272. Haqiqian, M.H.; Sardari, D.; Houshyari, M.; Moghadasali, R. Nano-sheets of Graphene Oxide Enhance the Combined Effect of Hyperthermia and Radiation Treatment in Pancreatic Cancer Cell Lines. Curr. Nanosci. 2021, 17, 779–788. [Google Scholar] [CrossRef]
  273. Wójcik, B.; Sawosz, E.; Szczepaniak, J.; Strojny, B.; Sosnowska, M.; Daniluk, K.; Zielińska-Górska, M.; Bałaban, J.; Chwalibog, A.; Wierzbicki, M. Effects of Metallic and Carbon-Based Nanomaterials on Human Pancreatic Cancer Cell Lines AsPC-1 and BxPC-3. Int. J. Mol. Sci. 2021, 22, 12100. [Google Scholar] [CrossRef]
  274. Caputo, D.; Coppola, A.; Quagliarini, E.; Di Santo, R.; Capriotti, A.L.; Cammarata, R.; Laganà, A.; Papi, M.; Digiacomo, L.; Coppola, R.; et al. Multiplexed Detection of Pancreatic Cancer by Combining a Nanoparticle-Enabled Blood Test and Plasma Levels of Acute-Phase Proteins. Cancers 2022, 14, 4658. [Google Scholar] [CrossRef] [PubMed]
  275. Caputo, D.; Digiacomo, L.; Cascone, C.; Pozzi, D.; Palchetti, S.; Di Santo, R.; Quagliarini, E.; Coppola, R.; Mahmoudi, M.; Caracciolo, G. Synergistic Analysis of Protein Corona and Haemoglobin Levels Detects Pancreatic Cancer. Cancers 2020, 13, 93. [Google Scholar] [CrossRef]
  276. Di Santo, R.; Digiacomo, L.; Quagliarini, E.; Capriotti, A.L.; Laganà, A.; Zenezini Chiozzi, R.; Caputo, D.; Cascone, C.; Coppola, R.; Pozzi, D.; et al. Personalized Graphene Oxide-Protein Corona in the Human Plasma of Pancreatic Cancer Patients. Front. Bioeng. Biotechnol. 2020, 8, 491. [Google Scholar] [CrossRef] [PubMed]
  277. Kang, S.; Kim, K.M.; Jung, K.; Son, Y.; Mhin, S.; Ryu, J.H.; Shim, K.B.; Lee, B.; Han, H.; Song, T. Graphene Oxide Quantum Dots Derived from Coal for Bioimaging: Facile and Green Approach. Sci. Rep. 2019, 9, 4101. [Google Scholar] [CrossRef] [PubMed]
  278. Nigam, P.; Waghmode, S.; Louis, M.; Wangnoo, S.; Chavan, P.; Sarkar, D. Graphene quantum dots conjugated albumin nanoparticles for targeted drug delivery and imaging of pancreatic cancer. J. Mater. Chem. B 2014, 2, 3190–3195. [Google Scholar] [CrossRef]
  279. Garriga, R.; Jurewicz, I.; Seyedin, S.; Bardi, N.; Totti, S.; Matta-Domjan, B.; Velliou, E.G.; Alkhorayef, M.A.; Cebolla, V.L.; Razal, J.M.; et al. Multifunctional, biocompatible and pH-responsive carbon nanotube- and graphene oxide/tectomer hybrid composites and coatings. Nanoscale 2017, 9, 7791–7804. [Google Scholar] [CrossRef] [PubMed]
  280. Ajgaonkar, R.; Lee, B.; Valimukhametova, A.; Nguyen, S.; Gonzalez-Rodriguez, R.; Coffer, J.; Akkaraju, G.R.; Naumov, A.V. Detection of Pancreatic Cancer miRNA with Biocompatible Nitrogen-Doped Graphene Quantum Dots. Materials 2022, 15, 5760. [Google Scholar] [CrossRef] [PubMed]
  281. Yoon, H.J.; Shanker, A.; Wang, Y.; Kozminsky, M.; Jin, Q.; Palanisamy, N.; Burness, M.L.; Azizi, E.; Simeone, D.M.; Wicha, M.S.; et al. Tunable Thermal-Sensitive Polymer-Graphene Oxide Composite for Efficient Capture and Release of Viable Circulating Tumor Cells. Adv. Mater. 2016, 28, 4891–4897. [Google Scholar] [CrossRef]
  282. Du, X.; Zheng, X.; Zhang, Z.; Wu, X.; Sun, L.; Zhou, J.; Liu, M. A Label-Free Electrochemical Immunosensor for Detection of the Tumor Marker CA242 Based on Reduced Graphene Oxide-Gold-Palladium Nanocomposite. Nanomaterials 2019, 9, 1335. [Google Scholar] [CrossRef]
  283. Wu, J.; Li, Z.; Li, Y.; Pettitt, A.; Zhou, F. Photothermal Effects of Reduced Graphene Oxide on Pancreatic Cancer. Technol. Cancer Res. Treat. 2018, 17, 1533034618768637. [Google Scholar] [CrossRef] [PubMed]
  284. Yu, Y.; Liang, C.; Wan, Q.Q.; Jin, D.; Liu, X.; Zhang, Z.; Sun, Z.Y.; Zhang, G.J. Integrated FET sensing microsystem for specific detection of pancreatic cancer exosomal miRNA10b. Anal. Chim. Acta 2023, 1284, 341995. [Google Scholar] [CrossRef] [PubMed]
  285. Akin, M.; Bekmezci, M.; Bayat, R.; Coguplugil, Z.K.; Sen, F.; Karimi, F.; Karimi-Maleh, H. Mobile device integrated graphene oxide quantum dots based electrochemical biosensor design for detection of miR-141 as a pancreatic cancer biomarker. Electrochim. Acta 2022, 435, 141390. [Google Scholar] [CrossRef]
  286. Xiao, G.; Ge, H.; Yang, Q.; Zhang, Z.; Cheng, L.; Cao, S.; Ji, J.; Zhang, J.; Yue, Z. Light-addressable photoelectrochemical sensors for multichannel detections of GPC1, CEA and GSH and its applications in early diagnosis of pancreatic cancer. Sens. Actuators B Chem. 2022, 372, 132663. [Google Scholar] [CrossRef]
  287. Chiu, N.-F.; Lin, T.-L.; Kuo, C.-T. Highly sensitive carboxyl-graphene oxide-based surface plasmon resonance immunosensor for the detection of lung cancer for cytokeratin 19 biomarker in human plasma. Sens. Actuators B Chem. 2018, 265, 264–272. [Google Scholar] [CrossRef]
  288. Gieseck Iii, R.L.; Hannan, N.R.F.; Bort, R.; Hanley, N.A.; Drake, R.A.L.; Cameron, G.W.W.; Wynn, T.A.; Vallier, L. Maturation of Induced Pluripotent Stem Cell Derived Hepatocytes by 3D-Culture. PLoS ONE 2014, 9, e86372. [Google Scholar] [CrossRef] [PubMed]
  289. Georgakopoulos, N.; Prior, N.; Angres, B.; Mastrogiovanni, G.; Cagan, A.; Harrison, D.; Hindley, C.J.; Arnes-Benito, R.; Liau, S.-S.; Curd, A.; et al. Long-term expansion, genomic stability and in vivo safety of adult human pancreas organoids. BMC Dev. Biol. 2020, 20, 4. [Google Scholar] [CrossRef] [PubMed]
  290. Sachs, N.; Clevers, H. Organoid cultures for the analysis of cancer phenotypes. Curr. Opin. Genet. Dev. 2014, 24, 68–73. [Google Scholar] [CrossRef]
  291. Romero-Calvo, I.; Weber, C.R.; Ray, M.; Brown, M.; Kirby, K.; Nandi, R.K.; Long, T.M.; Sparrow, S.M.; Ugolkov, A.; Qiang, W.; et al. Human Organoids Share Structural and Genetic Features with Primary Pancreatic Adenocarcinoma Tumors. Mol. Cancer Res. 2019, 17, 70–83. [Google Scholar] [CrossRef]
  292. Usman, O.H.; Zhang, L.; Xie, G.; Kocher, H.M.; Hwang, C.-i.; Wang, Y.J.; Mallory, X.; Irianto, J. Genomic heterogeneity in pancreatic cancer organoids and its stability with culture. NPJ Genom. Med. 2022, 7, 71. [Google Scholar] [CrossRef]
  293. Lai Benjamin, F.L.; Lu Rick, X.; Hu, Y.; Davenport, H.L.; Dou, W.; Wang, E.Y.; Radulovich, N.; Tsao, M.S.; Sun, Y.; Radisic, M. Recapitulating pancreatic tumor microenvironment through synergistic use of patient organoids and organ-on-a-chip vasculature. Adv. Funct. Mater. 2020, 30, 2000545. [Google Scholar] [CrossRef] [PubMed]
  294. Martinez Paino, I.M.; Santos, F.; Zucolotto, V. Biocompatibility and toxicology effects of graphene oxide in cancer, normal, and primary immune cells. J. Biomed. Mater. Res. A 2017, 105, 728–736. [Google Scholar] [CrossRef] [PubMed]
  295. Shams, M.; Guiney, L.M.; Huang, L.; Ramesh, M.; Yang, X.; Hersam, M.C.; Chowdhury, I. Influence of functional groups on the degradation of graphene oxide nanomaterials. Environ. Sci. Nano 2019, 6, 2203–2214. [Google Scholar] [CrossRef]
  296. Kumar, S.; Kumar, S.; Srivastava, S.; Yadav, B.K.; Lee, S.H.; Sharma, J.G.; Doval, D.C.; Malhotra, B.D. Reduced graphene oxide modified smart conducting paper for cancer biosensor. Biosens. Bioelectron. 2015, 73, 114–122. [Google Scholar] [CrossRef] [PubMed]
  297. Kang, S.; Lee, J.; Ryu, S.; Kwon, Y.; Kim, K.-H.; Jeong, D.H.; Paik, S.R.; Kim, B.-S. Gold Nanoparticle/Graphene Oxide Hybrid Sheets Attached on Mesenchymal Stem Cells for Effective Photothermal Cancer Therapy. Chem. Mater. 2017, 29, 3461–3476. [Google Scholar] [CrossRef]
  298. Neklyudov, V.V.; Khafizov, N.R.; Sedov, I.A.; Dimiev, A.M. New insights into the solubility of graphene oxide in water and alcohols. Phys. Chem. Chem. Phys. 2017, 19, 17000–17008. [Google Scholar] [CrossRef] [PubMed]
  299. Ali, J.; Li, Y.; Shang, E.; Wang, X.; Zhao, J.; Mohiuddin, M.; Xia, X. Aggregation of graphene oxide and its environmental implications in the aquatic environment. Chin. Chem. Lett. 2023, 34, 107327. [Google Scholar] [CrossRef]
  300. Ghulam, A.N.; Dos Santos, O.A.; Hazeem, L.; Pizzorno Backx, B.; Bououdina, M.; Bellucci, S. Graphene oxide (GO) materials—Applications and toxicity on living organisms and environment. J. Funct. Biomater. 2022, 13, 77. [Google Scholar] [CrossRef]
  301. Tiwari, S.K.; Mishra, R.K.; Ha, S.K.; Huczko, A. Evolution of graphene oxide and graphene: From imagination to industrialization. ChemNanoMat 2018, 4, 598–620. [Google Scholar] [CrossRef]
  302. Cho, H.; Kim, S.-m.; Liang, H.; Kim, S. Electric-potential-induced uniformity in graphene oxide deposition on porous alumina substrates. Ceram. Int. 2020, 46, 14828–14839. [Google Scholar] [CrossRef]
  303. Huch, M.; Bonfanti, P.; Boj, S.F.; Sato, T.; Loomans, C.J.; Van De Wetering, M.; Sojoodi, M.; Li, V.S.; Schuijers, J.; Gracanin, A. Unlimited in vitro expansion of adult bi-potent pancreas progenitors through the Lgr5/R-spondin axis. EMBO J. 2013, 32, 2708–2721. [Google Scholar] [CrossRef]
  304. Ren, Y.; Yang, X.; Ma, Z.; Sun, X.; Zhang, Y.; Li, W.; Yang, H.; Qiang, L.; Yang, Z.; Liu, Y.; et al. Developments and Opportunities for 3D Bioprinted Organoids. Int. J. Bioprint. 2021, 7, 364. [Google Scholar] [CrossRef]
  305. Zheng, F.; Xiao, Y.; Liu, H.; Fan, Y.; Dao, M. Patient-Specific Organoid and Organ-on-a-Chip: 3D Cell-Culture Meets 3D Printing and Numerical Simulation. Adv. Biol. 2021, 5, 2000024. [Google Scholar] [CrossRef] [PubMed]
  306. Rawal, P.; Tripathi, D.M.; Ramakrishna, S.; Kaur, S. Prospects for 3D bioprinting of organoids. Bio-Des. Manuf. 2021, 4, 627–640. [Google Scholar] [CrossRef]
  307. Hirt, C.K.; Booij, T.H.; Grob, L.; Simmler, P.; Toussaint, N.C.; Keller, D.; Taube, D.; Ludwig, V.; Goryachkin, A.; Pauli, C. Drug screening and genome editing in human pancreatic cancer organoids identifies drug-gene interactions and candidates for off-label therapy. Cell Genom. 2022, 2, 100095. [Google Scholar] [CrossRef] [PubMed]
  308. Liu, Y.; Wu, W.; Wang, Y.; Han, S.; Yuan, Y.; Huang, J.; Shuai, X.; Peng, Z. Recent development of gene therapy for pancreatic cancer using non-viral nanovectors. Biomater. Sci. 2021, 9, 6673–6690. [Google Scholar] [CrossRef] [PubMed]
  309. Cui, L.; Chen, Z.; Zhu, Z.; Lin, X.; Chen, X.; Yang, C.J. Stabilization of ssRNA on graphene oxide surface: An effective way to design highly robust RNA probes. Anal. Chem. 2013, 85, 2269–2275. [Google Scholar] [CrossRef]
  310. Vincent, M.; De Lázaro, I.; Kostarelos, K. Graphene materials as 2D non-viral gene transfer vector platforms. Gene Ther. 2017, 24, 123–132. [Google Scholar] [CrossRef] [PubMed]
  311. Beelen, N.A.; Aberle, M.R.; Bruno, V.; Olde Damink, S.W.M.; Bos, G.M.J.; Rensen, S.S.; Wieten, L. Antibody-dependent cellular cytotoxicity-inducing antibodies enhance the natural killer cell anti-cancer response against patient-derived pancreatic cancer organoids. Front. Immunol. 2023, 14, 1133796. [Google Scholar] [CrossRef]
  312. Liu, W.; Zhang, X.; Zhou, L.; Shang, L.; Su, Z. Reduced graphene oxide (rGO) hybridized hydrogel as a near-infrared (NIR)/pH dual-responsive platform for combined chemo-photothermal therapy. J. Colloid Interface Sci. 2019, 536, 160–170. [Google Scholar] [CrossRef]
  313. Costa, V.C.d.S. Preclinical Testing of Theranostic Graphene-Based Magnetic Nanocarriers in 2D and 3D Hepatocellular Carcinoma Models; University of Minho: Braga, Portugal, 2021. [Google Scholar]
  314. Darrigues, E.; Nima, Z.A.; Griffin, R.J.; Anderson, J.M.; Biris, A.S.; Rodriguez, A. 3D cultures for modeling nanomaterial-based photothermal therapy. Nanoscale Horiz. 2020, 5, 400–430. [Google Scholar] [CrossRef] [PubMed]
  315. You, M.; Zhu, H.; Li, Z.; Ye, E. Photothermal Nanomaterials for Oncological Hyperthermia. In Photothermal Nanomaterials; The Royal Society of Chemistry: London, UK, 2022; pp. 321–333. [Google Scholar]
  316. Zhou, C.; Wu, H.; Wang, M.; Huang, C.; Yang, D.; Jia, N. Functionalized graphene oxide/Fe3O4 hybrids for cellular magnetic resonance imaging and fluorescence labeling. Mater. Sci. Eng. C 2017, 78, 817–825. [Google Scholar] [CrossRef] [PubMed]
  317. Thakkar, S.; Sharma, D.; Kalia, K.; Tekade, R.K. Tumor microenvironment targeted nanotherapeutics for cancer therapy and diagnosis: A review. Acta Biomater. 2020, 101, 43–68. [Google Scholar] [CrossRef] [PubMed]
  318. Barar, J.; Omidi, Y. Dysregulated pH in tumor microenvironment checkmates cancer therapy. BioImpacts BI 2013, 3, 149. [Google Scholar]
Figure 1. Comprehensive overview of organoid development and tumor microenvironment modeling. The diverse origins of organoids, including resected tissues, biopsies, pluripotent stem cells, organ-specific adult stem cells, and 3D printing techniques are highlighted. The composition of the extracellular matrix (ECM) distinguishes between animal-derived and engineered components. In the context of the tumor microenvironment, the figure showcases the integration of organoids with elements such as culture medium, non-neoplastic cells, and angiogenesis, emphasizing the utility of organoids in replicating complex biological interactions. (Epidermal growth factor, EGF; Fibroblast growth factor 10, FGF10; Mouse Noggin Recombinant Protein, mNoggin; A83-01, Selective inhibitor of TGF-βRI, ALK4, and ALK7).
Figure 1. Comprehensive overview of organoid development and tumor microenvironment modeling. The diverse origins of organoids, including resected tissues, biopsies, pluripotent stem cells, organ-specific adult stem cells, and 3D printing techniques are highlighted. The composition of the extracellular matrix (ECM) distinguishes between animal-derived and engineered components. In the context of the tumor microenvironment, the figure showcases the integration of organoids with elements such as culture medium, non-neoplastic cells, and angiogenesis, emphasizing the utility of organoids in replicating complex biological interactions. (Epidermal growth factor, EGF; Fibroblast growth factor 10, FGF10; Mouse Noggin Recombinant Protein, mNoggin; A83-01, Selective inhibitor of TGF-βRI, ALK4, and ALK7).
Ijms 25 01066 g001
Figure 2. Schematic overview of graphene oxide applications in cancer diagnosis and treatment. Graphene oxide is widely utilized for its unique physical and chemical properties. Surrounding this central structure of GO are various methodologies employed in cancer diagnostics and therapeutics. On the left, diagnostic techniques are showcased: magnetic resonance imaging (MRI), fluorescence imaging, photoacoustic imaging, Raman imaging, and computed tomography (CT), highlighting the multifaceted applications of graphene oxide in enhancing imaging modalities. On the right, treatment approaches are depicted: drug delivery, gene delivery, antibody delivery, photothermal therapy, and photodynamic therapy, indicating the versatility of graphene oxide as a carrier and its role in targeted treatment strategies.
Figure 2. Schematic overview of graphene oxide applications in cancer diagnosis and treatment. Graphene oxide is widely utilized for its unique physical and chemical properties. Surrounding this central structure of GO are various methodologies employed in cancer diagnostics and therapeutics. On the left, diagnostic techniques are showcased: magnetic resonance imaging (MRI), fluorescence imaging, photoacoustic imaging, Raman imaging, and computed tomography (CT), highlighting the multifaceted applications of graphene oxide in enhancing imaging modalities. On the right, treatment approaches are depicted: drug delivery, gene delivery, antibody delivery, photothermal therapy, and photodynamic therapy, indicating the versatility of graphene oxide as a carrier and its role in targeted treatment strategies.
Ijms 25 01066 g002
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

Mai, S.; Inkielewicz-Stepniak, I. Graphene Oxide Nanoparticles and Organoids: A Prospective Advanced Model for Pancreatic Cancer Research. Int. J. Mol. Sci. 2024, 25, 1066. https://doi.org/10.3390/ijms25021066

AMA Style

Mai S, Inkielewicz-Stepniak I. Graphene Oxide Nanoparticles and Organoids: A Prospective Advanced Model for Pancreatic Cancer Research. International Journal of Molecular Sciences. 2024; 25(2):1066. https://doi.org/10.3390/ijms25021066

Chicago/Turabian Style

Mai, Shaoshan, and Iwona Inkielewicz-Stepniak. 2024. "Graphene Oxide Nanoparticles and Organoids: A Prospective Advanced Model for Pancreatic Cancer Research" International Journal of Molecular Sciences 25, no. 2: 1066. https://doi.org/10.3390/ijms25021066

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