Next Article in Journal
Photoacoustic/Ultrasound Endoscopic Imaging Reconstruction Algorithm Based on the Approximate Gaussian Acoustic Field
Next Article in Special Issue
Ratiometric Fluorescence Detection of Colorectal Cancer-Associated Exosomal miR-92a-3p with DSN-Assisted Signal Amplification by a MWCNTs@Au NCs Nanoplatform
Previous Article in Journal
Fluorescent Biosensors for the Detection of Viruses Using Graphene and Two-Dimensional Carbon Nanomaterials
Previous Article in Special Issue
Potentiometric Performance of Ion-Selective Electrodes Based on Polyaniline and Chelating Agents: Detection of Fe2+ or Fe3+ Ions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in Inflammatory Diagnosis with Graphene Quantum Dots Enhanced SERS Detection

by
Seyyed Mojtaba Mousavi
1,*,
Seyyed Alireza Hashemi
2,
Masoomeh Yari Kalashgrani
3,
Darwin Kurniawan
1,
Ahmad Gholami
3,
Vahid Rahmanian
4,
Navid Omidifar
5 and
Wei-Hung Chiang
1,*
1
Department of Chemical Engineering, National Taiwan University of Science and Technology, Taipei City 106335, Taiwan
2
Nanomaterials and Polymer Nanocomposites Laboratory, School of Engineering, University of British Columbia, Kelowna, BC V1V 1V7, Canada
3
Biotechnology Research Center, Shiraz University of Medical Science, Shiraz 71468-64685, Iran
4
Centre of Molecular and Macromolecular Studies, Polish Academy of Sciences, Sienkiewicza 112, 90-363 Lodz, Poland
5
Department of Pathology, School of Medicine, Shiraz University of Medical Sciences, Shiraz 71468-64685, Iran
*
Authors to whom correspondence should be addressed.
Biosensors 2022, 12(7), 461; https://doi.org/10.3390/bios12070461
Submission received: 28 May 2022 / Revised: 20 June 2022 / Accepted: 21 June 2022 / Published: 27 June 2022
(This article belongs to the Special Issue Electrochemistry and Spectroscopy-Based Biosensors)

Abstract

:
Inflammatory diseases are some of the most common diseases in different parts of the world. So far, most attention has been paid to the role of environmental factors in the inflammatory process. The diagnosis of inflammatory changes is an important goal for the timely diagnosis and treatment of various metastatic, autoimmune, and infectious diseases. Graphene quantum dots (GQDs) can be used for the diagnosis of inflammation due to their excellent properties, such as high biocompatibility, low toxicity, high stability, and specific surface area. Additionally, surface-enhanced Raman spectroscopy (SERS) allows the very sensitive structural detection of analytes at low concentrations by amplifying electromagnetic fields generated by the excitation of localized surface plasmons. In recent years, the use of graphene quantum dots amplified by SERS has increased for the diagnosis of inflammation. The known advantages of graphene quantum dots SERS include non-destructive analysis methods, sensitivity and specificity, and the generation of narrow spectral bands characteristic of the molecular components present, which have led to their increased application. In this article, we review recent advances in the diagnosis of inflammation using graphene quantum dots and their improved detection of SERS. In this review study, the graphene quantum dots synthesis method, bioactivation method, inflammatory biomarkers, plasma synthesis of GQDs and SERS GQD are investigated. Finally, the detection mechanisms of SERS and the detection of inflammation are presented.

1. Introduction

When tissue is damaged by bacteria, trauma, chemicals, heat, or other phenomena, several substances are released from the damaged tissue that cause very serious secondary changes in the tissue. This series of tissue changes is called inflammation. Inflammation is basically a protective response that fights the cause of cellular damage (e.g., germs or toxins) and the consequences of that damage, i.e., necrotic cells and tissues [1,2,3,4]. Inflammation causes hypersensitivity reactions to insect bites, drugs, and toxins, as well as some chronic diseases, such as rheumatoid arthritis, atherosclerosis, and pulmonary fibrosis [5,6,7,8]. Inflammation is a complex process that begins with tissue damage caused by endogenous factors, such as tissue necrosis and bone fractures, or exogenous factors, such as mechanical, physical, biological damage (like infection with microorganisms or immunological responses, such as hypersensitivity reactions), and is accompanied by the invasion of inflammatory cells into the inflamed area [9,10]. Graphene quantum dots are a new generation of quantum dot structures. This new generation of carbon compounds contains a large number of functional groups and is synthesised in very small dimensions that exhibit a variety of other properties of carbon-based materials. Graphene quantum dots generally have a sheet structure of less than 10 nm in size and have received much attention in recent years due to their unique chemical and physical properties [11,12]. Compared with semiconductor quantum dots, graphene quantum dots are characterised by high water solubility, good photoluminescence properties, biocompatibility, optimal accessibility, easy surface functionalization, high stability, and low toxicity [13,14]. Thus, GQDs effectively reduce hyperinflammation by regulating immune cells, suggesting that they can be used as promising diagnostic agents to diagnose inflammation [15,16]. Surface-enhanced Raman spectroscopy (SERS) has also attracted great interest in various fields, such as medicine and analytical chemistry, due to its unique properties—sensitivity to single molecules on the surface, multiplexing potential, and fingerprinting capabilities [17,18,19,20]. One of the most important substrates of SERS, which includes quantum dots, has a wide range of applications in inflammation diagnosis, biological and chemical imaging, and labelling due to the plasmonic resonance properties of the surface-dependent local area, the chemical richness of the core-shell, and the compatible size [21,22,23,24,25]. In addition, graphene quantum dots also serve as a building block for an atomically flat SERS substrate, in which a much more uniform Raman signal can be obtained [26,27,28]. Graphene quantum dots (GQDs) have more accessible edges and larger specific surface areas than conventional graphene sheets, resulting in more efficient adsorption of target molecules [29,30,31].
This review study aims to explore recent advances in inflammation diagnostics using graphene quantum dots to improve the detection of SERS and highlight future areas of work in this field. In addition, the graphene quantum dots, bioactivation method, GQD synthesis method, inflammation biomarkers, plasma synthesis of GQDs, and SERS GQD were investigated. In addition, the detection methods of SERS and the detection of pro-inflammatory substances were evaluated.

2. Graphene Quantum Dot

Graphene quantum dots have attracted the attention of many researchers due to the crystalline structure of a single or a small amount of crushed graphene. These nanoparticles are a small lattice structure of honeycombs of carbon atoms that are less than 10 nanometers in size. Graphene quantum dots (GQDs) are, by definition, a type of quantum dot material with a graphene-derived property and carbon dots that can be placed in the form of very small graphene pieces (Figure 1). The carbon dots exhibit strong optical absorption in the UV region (260–320 nm) with an extended sequence in the visible and infrared regions. In addition, GQDs are semiconductor quantum dots with excellent light stability, biocompatibility and low toxicity, good electron mobility and good chemical stability, small size, electrochemical luminescence, photocatalyst capabilities, and are suitable for the fabrication of multiple sensors and bioimaging [32,33,34,35,36]. GQDs are less toxic than graphene oxides and have no obvious toxicity in the body, so GQDs have attracted much attention in biological applications, especially in the field of biopharmaceuticals [37,38]. Numerous groups have explored recent advances in graphene quantum dots for the construction of various sensors, including electron sensors, photoluminescence sensors (PL), electrochemical sensors, electrochemical luminescence sensors, PL-based high-conversion sensors, and surface-enhanced Raman spectroscopy (SERS) [39,40,41,42]. GQDs have become a prominent substance for the design of SERS due to their outstanding properties and model assumptions, such as high electron transfer rate, higher biomolecule loading, fast transduction, larger surface areas, easy surface functionalization, and inimitable electrocatalytic properties. These GQD enhanced SERS have been utilised for the detection of nucleic acids, amino acids, bioflavonoids, vitamins, small molecules, biomarkers, and heavy metal ions with remarkable properties [43].

2.1. Method of Synthesis GQD

Graphene quantum dots are synthesized to optimize the size of quantum dots by two methods: top-down and bottom-up (Figure 2). In the top-down method, bulky carbon, graphite, and graphene materials are converted into graphene quantum dots, while in the bottom-up method, organic molecules are used as the carbon source. The disadvantages of the top-down method are the difficulties in controlling the size distribution and morphology of the produced particles. In contrast, the properties of the produced nanoparticles can be well controlled by the bottom-up method. In general, the optical properties of graphene nanodots depend on the size and the effect of quantum confinement, which changes the density and the nature of sp2 sites. Therefore, the energy of these nanoparticles changes with the size of the gap [44,45,46,47]. Functionalizing the surface and doping the GQDs with other elements are other possible strategies to change these energy gaps while increasing the photoluminescence quantum yield (PLQY) of the GQDs by suppressing the emitting traps.
In the top-down method, coal, which is considered the cheapest and easiest material to cleave compared to the other available materials, is exfoliated to form GQDs. For example, Ye et al. first sonicated coal in a mixture of concentrated sulfuric acid and nitrile acid for 2 h before heat-treating the mixture in an oil bath at 100–120 °C for 24 h to produce GQDs [48]. Yan et al. succeeded in controlling the band gap of the coal-derived GQDs using a surface functionalization technique. In the typical procedure, the coal-derived GQDs were mixed in a toluene solution with various organic compounds (e.g., o-phenylenediamine, 2,3-diaminonaphthalene, 1,8-diaminonaphthalene, 1,1′-bi(2-naphthylamine), p-anisidine, 4-(trifluoromethoxy)-aniline, or 4-(trichloromethoxy)-aniline) and then solvothermally treated at 180 °C for 12 h to systematically adjust the band gap of the GQDs [49]. Since there are many concerns about the use of strong concentrated acids, Shin et al. prepared GQDs from various natural carbon sources using an acid-free oxone oxidant-assisted solvothermal technique [50]. Another acid-free strategy based on the ultrasonic irradiation of a mixture of anthracite charcoal and N, N-dimethylformamide (DMF) was used by Zhang et al. to prepare GQDs [51]. Since most of the feedstocks used are non-renewable sources and sometimes special chemicals are required to obtain GQDs with tunable emissions, and since high temperatures and long reaction times are also required, the feasibility of the top-down method is significantly hindered, especially when addressing the current problem of global energy limitations.
In contrast, the bottom-up method uses polycyclic aromatic compounds or other molecules with an aromatic structure, such as fluorene [25,52,53,54,55]. Table 1 shows the characteristics of the top-down and bottom-up methods in the synthesis of GQDs. However, the toxicity of these aromatic precursors may have a negative impact on the environment and is therefore considered unsuitable for large-scale production. For this reason, many efforts have been made to utilize naturally available biomasses as the main starting materials for the synthesis of GQDs. Citric acid, as one of the most commonly used biomasses, can be pyrolyzed directly at 200 °C to obtain blue-emitting GQDs [56]. Nitrogen-containing molecules can be used, together with citric acid, to prepare readily nitrogen-doped GQDs (NGQDs). Wu et al. synthesized blue-emitting NGQDs with a PLQY of 36.5% from a mixture of citric acid and dicyandiamide using a hydrothermal technique at 180 °C for 3 h [57]. Recently, biomass waste has attracted much attention due to its low cost, renewable and environmentally friendly properties. Kumar et al. reported a one-step preparation of NGQDs from chitosan using a chemical vapor deposition (CVD) system at 250–300 °C [58]. To avoid the use of high temperatures, Chiang’s group used microplasma technologies to synthesize colloidal NGQDs with a PLQY of 30% from chitosan at ambient conditions [59]. Various strategies involving plasma flow, reaction time, and the type of acid used to dissolve chitosan were employed to control the functionalities and thus the energy gap of the resulting NGQDs [59,60]. Another promising biomass waste as a GQD precursor is lignin, which consists of phenyl skeletons and oxygenated branches [61]. Unlike simple structures, such as citric acid and glucose, both chitosan and lignin are biopolymers with complex structures, so the synthesis mechanism could involve a combination of top-down and bottom-up processes. The underlying mechanism is thought to consist of two main steps. These include the decomposition of long-chain structures into smaller units, the subsequent refusion into a nanograph domain, and the growth of GQDs [61,62]. Overall, the possibilities of the bottom-up method to utilize biomass derivatives as GQDs precursors have gained much interest nowadays, in order to realize a more sustainable, green, and eco-friendly approach to synthesize GQDs with unique properties and controlled structures to be usable for many applications.

2.2. Plasma Synthesis of GQDs

Plasma synthesis is considered one of the most popular gas-phase methods for the preparation of various GQDs, especially those with covalent bonding [70,71,72]. For example, QDs of germanium (Ge) and silicon (Si) have been synthesized using a conventional non-thermal plasma. In non-thermal plasma, factors, such as shape, surface area, quantum dot composition and size, can be controlled [73,74]. Plasma synthesis achieves doping, which is a major challenge for QDs [75,76,77]. GQDs synthesized by plasma usually take the form of powder, which can lead to surface modification. This can lead to excellent dispersion of QDs in water [78] or organic solvents [79] (i.e., colloidal quantum dots).

2.3. Method of Bioactivation

2.3.1. Bioactive Carbon Sources

The development of bioactive materials for biomedical applications, such as inflammation therapy, is desirable if it is compatible with detectable properties and integrates efficient differentiation into biocompatible procedures. It has been possible to fabricate bioactive carbon dots (CD) with a size of about 4 nm, which have low toxicity, interesting safety responses, and unique photophysical properties. Bioactive CDs were prepared by a novel one-step hydrothermal method from aspirin and adenosine [80,81,82,83]. Multipurpose CDs are designed and fabricated using a bottom-up synthesis strategy to further manipulate chemical compounds and physical properties by introducing complex bioactive precursors, including nucleic acids, proteins, and small molecules. These bioactive CDs have different pharmacological activity from conventional citric acid-based CDs to expand their potential applications against pathogens and cancer [84,85].

2.3.2. Biomass-Waste Derived GQD

Due to increasing customer demand, energy crisis and environmental degradation, scientists are looking for cost-effective, environmentally friendly, and easy ways to produce new advanced materials from renewable sources. Recently, graphene quantum dots (GQDs) have attracted much attention compared to other investigated materials, such as carbon-based nanomaterials, due to their attractive properties, such as low toxicity, long lifetime, high conductivity, good biocompatibility, and large surface area. Therefore, the properties of biomass waste-derived GQDs have been modified by adding surface inactivating agents and various functional groups through surface processing [31,86]. Kalita et al. investigated the modification of GQDs by amine functionalization to enhance the quantum efficiency of rice grain-derived GQDs. The results showed that the quantum efficiency was improved by 125% after amine functionalization, which was attributed to the superior electron donating ability of amine groups [31,86]. Table 2 shows the synthesis of GQD from different types of biomass waste. Figure 3a,b show GQDs obtained from biowaste and different approaches to convert biowaste into GQDs, with a description of their use as energy sources.

2.3.3. Biologically Active Agents

Bioactive agents are factors that affect a living organ, cell, or tissue. They may be bioactive compounds, vitamins, drugs, phytochemicals, or enzymes. Bioactive agents used in biomedical devices and drugs can be contained in polymers [92]. The loading of bioactive agents in drug delivery systems is carried out by enzymatically reacting polymers and the cleavage of these agents by the target enzymes. The release of the therapeutic cargo occurs through the activation of the bioactive agents [93].

3. SERS GQD

The frequency of the Raman peak of a phonon can be related to the chemical composition, the internal stress or surface state, and the shape and size of the GQD. Applications of Raman spectroscopy include the multiplexed detection of biomarkers from various compounds, controlling the synthesis of GQDs, studying vibrational properties associated with relaxation mechanisms, and analysing GQDs. For example, Raman spectra are associated with different concentrations of alloying constituents for alloyed GQDs, and this spectrum is also used to indicate the formation of quantum dots [94,95]. Therefore, the use of surface-enhanced Raman spectroscopy (SERS) is possible when the Raman intensity is undetectable and very low, such as the deposition of nanoparticles on a metal film, such as Ag, Al, or Au; these films have surface plasmon states with very limited electromagnetic fields that support emission and light absorption at the interface between air and metal. SERS also has applications from vibrational spectroscopy or photochemical studies [96] to the single emitter level [97,98], such as the detection of bacteria or impurities, and chemical analysis. Since different fabrication methods and many substrates have been proposed, it can be said that in the search for efficient and cheap metal substrates, SERS is a very active field. Examples are the gold nanoparticles recently obtained by the femtosecond exposure of a gold film [99], the strips of vertical or horizontal gold nanorods [100], the dual pyramidal gold nanoparticles [101], or the growth of silver nanoparticles on Si by the alternative method [100]. Hot spots occur when the key element is the presence of sharp points or dents on the surface of the metal substrate, where plasmonic states can be strongly confined. Hot spots are also crucial for future plasmonic applications in photochemistry or photovoltaics [96]. Some works have applied SERS to GQDs made of different materials: SnO2 [102], Si [103], CdS [104], PbS [105], or CdSe [106,107,108]. Thus, SERS is used as a promising technique for GQD detection; biomarker detection; in situ tracking of nanoparticle surface chemical properties, such as oxidation; and nanoparticle impact detection [26,103,105]. SERS in GQDs can be very useful because the material is not only environmentally friendly and compatible with the SERS mechanism but also has a large specific surface area, biocompatibility, and high chemical stability [109], which supports the improvement process of the mechanism for the improved detection of GQDs SERS. Although previous work SERS has reported GQDs in solid form, compared to GQDs in solution-based formats, it may require a complex process to use GQDs in solid form for other SERS applications. In addition, solution-based GQDs can be synthesized by a top-down approach, such as an electrochemical method that provides the ability to obtain SERS of GQDs in solution. This method uses the electric field as a mechanism to support the enhanced SERS detection of GQDs in the chemical mechanism (CM) to initiate the contact process of molecules and chemical reactions [110]. A schematic representation of the mechanism of SERS GQDs for analyte detection can be found in Figure 4. Graphene and other 2D materials have been developed for use as Raman enhancement substrates. This is due to their unique single sheet of carbon atoms in a 2D honeycomb crystal structure of electrons and phonons with one 2pz orbital of each sp2 hybridized carbon atom constituting a large, delocalized bond, forming an ideal flat surface and strong chemical interaction with many organic molecules [111]. As a SERS platform, graphene thus enables the independent investigation of the chemical enhancement mechanism (CM). Graphene may boost the Raman signals of molecules that have been adsorbed, and these substrates have shown to be promising for the detection of micro- and trace species. This impact was identified for the first time in 2010 by Ling et al. [112]. When graphene is treated with organic solvents, several “emerging bands” form on mechanically exfoliated graphene [113]. These “emerging bands” are scattered among the unidentified organic compounds included in the transparent tape used for graphene exfoliation. Graphene has a Raman amplification effect on trace residues. GQDs SERS is useful for both the fundamental investigations of SERS phenomena and several practical applications because of the graphene matrix’s significant benefits, such as homogeneity, repeatability, cleanliness, and low detection limit for aromatherapy dyes. Graphene makes SERS applications of Raman-enhanced substrates more quantitatively controlled.

4. Detection Mechanisms of SERS

Since the discovery of SERS, several mechanisms have been proposed, but only two are widely accepted today: the electromagnetic theory (EM) and the chemical amplification theory (CE). The EM theory is more dominant because it can amplify the Raman signal up to ten thousand times. While CE amplifies the Raman signal up to 100 times. In the EM model, the laser interacts with the metal surface. As a result of this interaction, dissolved surface plasmons are stimulated, which amplify the field near the surface. In the CE theory, the electron states of the adsorbent change due to chemical adsorption. In the SERS phenomenon, both factors occur simultaneously, which is why the Raman signal can be amplified to such an extent. In general, surface-enhanced Raman spectroscopy is very similar to Raman resonance spectroscopy. The difference being that the resonances present are not exclusively of the intramolecular type. Surface-enhanced Raman spectroscopy or SERS is also a method of Raman spectroscopy that has very high sensitivity in deciphering materials [115,116,117]. This method has numerous applications in medical science. Since it does not damage natural tissues, it does not require sample preparation and is very rapid. Therefore, this method is used to detect proteins in body fluids. It is also used for diagnosis and treatment of tumours and cancer, the treatment of neurological diseases, the detection of COVID-19, and the detection of coronavirus RNA. This technology for detecting urea and label-free plasma in human serum can be useful in cancer screening. SERS can be used for drugs, forensics, the detection of drugs and explosives, the study of redox processes at the single molecule level, the quantitative analysis of small molecules in human biological fluids, the quantitative detection of biomolecular interactions and more. In a tumour testing, the tumour is grown in vitro. In reality, the test is performed on living tissue (these tests are called in vivo tests). SERS can be used to detect low molecular weight biomolecules (Figure 5) [118,119,120,121]. Moreover, the mechanism of SERS detection in the efficiency of GQDs due to surface factorization and heteroatomic doping has been discussed in few studies. Therefore, in the near future, in-depth studies on these topics, which open a new window to developing highly effective improved GQDs SERS, will help scientists and researchers to understand the inflammatory diagnosis of GQDs. Finally, low-cost industrial production is urgently needed to develop and expand these approaches in the near future. To diagnose the inflammation and interior of GQDs and control the macroscopic properties of GQDs, one must be able to produce GQDs by controlling their size as much as possible [122].

5. Inflammatory Biomarkers

The term biomarker was first used by the National Institutes of Health in the United States in 1980 [123]. Biomarkers are a new method in medicine in which these markers measure specific indicators and examine routine biological processes, pathological processes, or pharmacological responses through therapeutic or health mediators. Specific RNA/DNA gene sequences, antibody determinations, and organic metabolite measurements are also identified [124,125,126]. For example, blood pressure is a biomarker for stroke risk and glucose levels are a biomarker for patients with diabetes; cholesterol levels are also used to determine cardiovascular disease risk [127]. For the nervous system, muscle, blood, nerve, cerebrospinal fluid, skin, and urine have been used to extract information from the brain in healthy and unhealthy states. These tools and technologies directly measure biological factors (such as blood or CSF) or they work together with brain imaging to measure changes in the composition, function, and structure of the nervous system. Biomarkers are classified according to the sequence of events from “exposure” to “disease”. Biomarker “exposure” is used to predict hazards and to establish a link between external exposures and internal dosimetry. Disease biomarkers are used for the screening, identification, and monitoring of disease progression [128,129,130]. In addition, these biomarkers can be used to refine drugs to improve therapeutic outcome and health [127]. A good biomarker must be more than 80% specific and have an equally high sensitivity (above 80). The role of biomarkers is not only identification, but they also have the potential to be predictive or play a role in the development of new therapies [131]. Plasma biomarkers are divided into pro-inflammatory biomarkers that include the following subgroups [132]:

5.1. C-Reactive Protein and Cytokines

C-reactive protein (CPR) is a protein involved in host safety and is mainly released by adipose tissue and liver in response to inflammatory stress. On the other hand, its reaction with the crystallisable receptor fragment leads to the production of pre-inflammatory cytokines. According to studies, CRP levels increase in patients with migraines and in women who suffer migraines with aura [132,133,134]. Cytokines are small proteins that are released by the stimulating neuropeptides involved in migraines. Therefore, their serum levels increase during migraine attacks, for example [132]: Tumour necrosis factor alpha (TNF-α), as a proinflammatory cytokine, plays a key role in the regulation of immune cells; it is also involved in clot formation, cell proliferation, apoptosis, lipid metabolism, and increases in plasma after migraine attacks. Transforming growth factor beta 1 (TGF-β1), a proinflammatory cytokine, also has several functions. This type of cytokine not only plays an important role in immune system function and blood vessel formation, but also causes motility, apoptosis, cell growth control, and differentiation [132,134,135]. TGF-β1 levels are increased in migraine patients compared to controls [133], but there is no difference between levels in aura and without aura. Fatigue and lack of energy during migraines are due to increased TGF-β1 levels [136]. Figure 6 shows a schematic of C-reactive proteins and cytokines, the functional pathways of CRP. As a result of cytokines, such as IL6 and IL1*, hepatic CRP expression increases significantly. CRP circulates, opsonizing bacteria and apoptotic cells so they can be cleared through the complement system. Immunomodulatory cytokines, such as IL10, may be released by phagocytic cells in response to CRP ligation. Studies have found that plasma CRP deposited onto inflamed tissue breaks down into biologically active monomeric subunits, which can be credited with proinflammatory effects.

5.2. Adiponectin and Lipids

Adiponectin is released from adipose tissue and is an anti-inflammatory cytokine, like IL-10, that inhibits the expression of pre-inflammatory cytokines. It plays an important role in regulating glucose homeostasis and other metabolic processes and is associated with obesity and BMI [132,133]. Research shows that lipids are associated with high cholesterol and migraine. In addition to total cholesterol, there is evidence that people with migraine have further increases in lipid subtypes, such as low-density lipoprotein cholesterol, oxidised LDL-c, triglycerides, and also a decrease in the anti-inflammatory high-density lipoprotein cholesterol [132,137]. Figure 7 shows the main processes by which adiponectin maintains metabolic homeostasis.

5.3. Raman Spectrum of the Inflammatory Biomarkers

Recently, biomarker detection based on Raman spectrum technology has been widely and comprehensively developed. Raman spectrum technology has attracted more attention with the rapid development of nanotechnology. Raman spectrum technology plays an important diagnostic role from organelle functionality, inflammation detection, and virus detection to cell activity detection. However, there are still many challenges in the development of Raman spectrum technology [138,139]. Monitoring different types of inflammation in the early stages by in vitro diagnosis is vital. For early detection and prognosis of inflammation, biomarkers, such as proteins, miRNAs, DNAs, and other biomolecules, must be evaluated [140,141]. The specificity and sensitivity of the Raman spectrum make it possible to accurately detect related physiological analytes in complex biological fluids. To detect inflammatory biomarkers, the Raman spectrum is linked to relevant detection molecules (such as antibodies and aptamers) to allow specific targets to be measured with Raman signals [142,143].

6. Detection of Inflammatory

The immune system is involved in the development of a variety of diseases [144,145,146]. The regulation of immune responses by direct tissue imaging in diseases, such as atherosclerosis [147], rheumatoid arthritis [148], malaria [149], and stroke [150], is due to the increasing interest in understanding the molecular and cellular interactions of the pathway. A number of intravital microscopy techniques have been used to study these interactions. For example, the two-photon fluorescence microscope (TPM) has become the method of choice for stimulating fluorophores deep within tissues due to its unique ability to produce light in the near-infrared range. However, successful imaging is limited to depths of a few hundred micrometres [151]. In addition, fluorescence imaging produces a broad emission spectrum [152] that often leads to photobleaching [153], as this imaging generally suffers from poor discrimination between specific fluorophores and background autofluorescence beyond a small optical window [154]. One way to enhance the Raman signal is to use the SERS method. This method involves vibrational spectroscopy in which molecules are adsorbed onto a metal surface with a nano-sized surface area. SERS-activated GQDs were encoded with a unique Raman signal that was monitored under a wide range of excitations and conditions. GQDs containing active Raman molecules were conjugated with specific monoclonal antibodies against intercellular adhesion molecule 1 (ICAM-1) to detect early-stage inflammation. The non-invasive measurement of ICAM-1 expression by SERS is possible in vivo with double the sensitivity of double photon fluorescence [155]. Therefore, a new approach for the diagnosis of inflammation in vivo using GQDs SERS was considered. Using a metal surface to enhance Raman scattering from molecules located near or attached to the surface results in vibrational spectroscopy called SERS. A wide range of different surfaces and metals can be used to achieve this goal. GQDs, however, offer a great format. By binding molecules with a unique and strong Raman spectrum, called Raman reporters, to GQDs and encapsulating them in a silicon-containing shell, GQDs with improved SERS are produced. The Raman reporter molecules are protected by a silica shell that acts as a coating and gives the GQD a unique and strong SERS signal [156,157]. In vivo imaging of SERS-enriched GQDs has also been used to monitor inflammation and reuse. Although this is not a disease process, it is still important as any changes may indicate infection and non-hidden disease states Figure 8 [155].

7. Perspectives

The multiple applications of Raman and the significant improvement of GQDs SERS for the diagnosis of inflammatory diseases are direct consequences of the numerous advantages they offer. Unlike their fluorescent counterparts, they produce specific, sharp molecular spectra that give these techniques immediate and easy access to multiple ways of diagnosing disease. In addition, there is the possibility of combined Raman and SERS imaging of tissues and cells, so that biochemical information and features of the inflammatory process can be obtained simultaneously as active SERS nanotags are formed. To improve GQDs SERS and remain at the forefront of inflammatory disease diagnosis, further studies in physiological representative media, clinical samples, and in vivo are required. Once these programmes have fully demonstrated their reproducibility, sensitivity, robustness, repeatability and selectivity in vivo, the laboratory will be transferred to a clinical setting. Once these applications have fully demonstrated their techniques, sensitivity, and reliability they will be strong contenders that will revolutionise our ability to diagnose inflammatory diseases. Ultimately, the hope is that the use of GQDs, which SERS enhance for the in vivo diagnosis of inflammation, will be part of a growing toolkit for next-generation non-invasive imaging and in vivo diagnosis.

8. Conclusions

This review summarises recent advances in the diagnosis of inflammation using graphene quantum dot SERS. Three subtopics are described, including the method of GQD synthesis, the method of bioactivation, inflammatory biomarkers, the plasma synthesis of GQDs and SERS GQD, and the detection mechanisms of SERS and the detection of inflammation. There are key points in the development of SERS GQD and its biomedical applications, as the rapid evolution of SERS GQD from biological to biomedical applications has been remarkable over the past decades. Significant progress has been made in improving diagnostic sensitivity and multiplexity. Recent advances have led to SERS GQD being used to diagnose inflammation in necrotic tissue and damaged cells, which will be of great importance for use in medical facilities. In short, SERS GQD has the advantageous properties of unprecedented multiplexing capability, perfect signal specificity and high sensitivity. Therefore, there are driving forces to exploit these properties for important applications. However, there are still some steps to be taken when using SERS GQD for clinical applications.

Author Contributions

S.M.M. and A.G. developed the idea and structure of the review article. V.R., S.A.H. and M.Y.K. wrote the manuscript, collecting the materials from databases. N.O., D.K. and A.G. revised and improved the manuscript. A.G. and W.-H.C. supervised the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work is sponsored by the Ministry of Science and Technology, Taiwan (grant number: MOST 110-2628-E-011-003, MOST 109-2923-E-011-003-MY, MOST 111-NU-E-011-001-NU).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data generated or analysed during this study are included in thepublished article.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Khan, S.; Nisar, M.; Rehman, W.; Khan, R.; Nasir, F. Anti-inflammatory study on crude methanol extract and different fractions of Eremostachys laciniata. Pharm. Biol. 2010, 48, 1115–1118. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Stewart, F.; Akleyev, A.; Hauer-Jensen, M.; Hendry, J.; Kleiman, N.; Macvittie, T.; Aleman, B.; Edgar, A.; Mabuchi, K.; Muirhead, C. ICRP publication 118: ICRP statement on tissue reactions and early and late effects of radiation in normal tissues and organs–threshold doses for tissue reactions in a radiation protection context. Ann. ICRP 2012, 41, 1–322. [Google Scholar] [CrossRef] [PubMed]
  3. Kalashgarani, M.Y.; Babapoor, A. Application of nano-antibiotics in the diagnosis and treatment of infectious diseases. Adv. Appl. NanoBio-Technol. 2022, 3, 22–35. [Google Scholar]
  4. Mousavi, S.M.; Hashemi, S.A.; Amani, A.M.; Saed, H.; Jahandideh, S.; Mojoudi, F. Polyethylene terephthalate/acryl butadiene styrene copolymer incorporated with oak shell, potassium sorbate and egg shell nanoparticles for food packaging applications: Control of bacteria growth, physical and mechanical properties. Polym. Renew. Resour. 2017, 8, 177–196. [Google Scholar] [CrossRef]
  5. Fairweather, D. Atherosclerosis and Inflammatory Heart Disease. In Immunotoxicity, Immune Dysfunction, and Chronic Disease; Springer: Berlin/Heidelberg, Germany, 2012; pp. 271–289. [Google Scholar]
  6. Pascoal, A.; Estevinho, M.M.; Choupina, A.; Sousa-Pimenta, M.; Estevinho, L.M. An overview of the bioactive compounds, therapeutic properties and toxic effects of apitoxin. Food Chem. Toxicol. 2019, 134, 1–11. [Google Scholar] [CrossRef] [PubMed]
  7. Alipour, A.; Kalashgarani, M.Y. Nano Protein and Peptides for Drug Delivery and Anticancer Agents. Adv. Appl. NanoBio-Technol. 2022, 3, 60–64. [Google Scholar]
  8. Mousavi, S.M.; Hashemi, S.A.; Esmaeili, H.; Amani, A.M.; Mojoudi, F. Synthesis of Fe3O4 nanoparticles modified by oak shell for treatment of wastewater containing Ni (II). Acta Chim. Slov. 2018, 65, 750–756. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Mc-Namara, D.; Mayeux, P. Nonopiate analgesics and anti-inflammatory drugs. In Principles of Pharmacology: Basic Concepts and Clinical Applications, 3rd ed.; Int. Thomson Publ. Co., Chapman Hall: Pacific Grove, CA, USA, 1995; pp. 1160–1178. [Google Scholar]
  10. Mousavi, S.; Hashemi, S.; Gholami, A.; Kalashgarani, M.; Vijayakameswara Rao, N.; Omidifar, N.; Hsiao, W.; Lai, C.; Chiang, W. Plasma-Enabled Smart Nano Exosome Platform as Emerging Immunopathogenesis for Clinical Viral Infection. Pharmaceutics 2022, 14, 1054. [Google Scholar] [CrossRef]
  11. Xie, R.; Wang, Z.; Zhou, W.; Liu, Y.; Fan, L.; Li, Y.; Li, X. Graphene quantum dots as smart probes for biosensing. Anal. Methods 2016, 8, 4001–4016. [Google Scholar] [CrossRef]
  12. Kazemi, K.; Ghahramani, Y.; Kalashgrani, M.Y. Nano biofilms: An emerging biotechnology applications. Adv. Appl. NanoBio-Technol. 2022, 3, 8–15. [Google Scholar]
  13. Farshbaf, M.; Davaran, S.; Rahimi, F.; Annabi, N.; Salehi, R.; Akbarzadeh, A. Carbon quantum dots: Recent progresses on synthesis, surface modification and applications. Artif. Cells Nanomed. Biotechnol. 2018, 46, 1331–1348. [Google Scholar] [CrossRef]
  14. Mousavi, S.M.; Hashemi, S.A.; Kalashgrani, M.Y.; Gholami, A.; Omidifar, N.; Babapoor, A.; Vijayakameswara Rao, N.; Chiang, W.-H. Recent Advances in Plasma-Engineered Polymers for Biomarker-Based Viral Detection and Highly Multiplexed Analysis. Biosensors 2022, 12, 286. [Google Scholar] [CrossRef] [PubMed]
  15. Lee, B.-C.; Lee, J.Y.; Kim, J.; Yoo, J.M.; Kang, I.; Kim, J.-J.; Shin, N.; Kim, D.J.; Choi, S.W.; Kim, D. Graphene quantum dots as anti-inflammatory therapy for colitis. Sci. Adv. 2020, 6, eaaz2630. [Google Scholar] [CrossRef]
  16. Mousavi, S.; Arjmand, O.; Hashemi, S.; Banaei, N. Modification of the epoxy resin mechanical and thermal properties with silicon acrylate and montmorillonite nanoparticles. Polym. Renew. Resour. 2016, 7, 101–113. [Google Scholar] [CrossRef]
  17. Qian, X.-M.; Nie, S.M. Single-molecule and single-nanoparticle SERS: From fundamental mechanisms to biomedical applications. Chem. Soc. Rev. 2008, 37, 912–920. [Google Scholar] [CrossRef] [PubMed]
  18. Gubala, V.; Harris, L.F.; Ricco, A.J.; Tan, M.X.; Williams, D.E. Point of care diagnostics: Status and future. Anal. Chem. 2012, 84, 487–515. [Google Scholar] [CrossRef]
  19. Bonaccorso, F.; Colombo, L.; Yu, G.; Stoller, M.; Tozzini, V.; Ferrari, A.C.; Ruoff, R.S.; Pellegrini, V. Graphene, related two-dimensional crystals, and hybrid systems for energy conversion and storage. Science 2015, 347, 1246501. [Google Scholar] [CrossRef]
  20. Tian, L.; Tadepalli, S.; Fei, M.; Morrissey, J.J.; Kharasch, E.D.; Singamaneni, S. Off-resonant gold superstructures as ultrabright minimally invasive surface-enhanced Raman scattering (SERS) probes. Chem. Mater. 2015, 27, 5678–5684. [Google Scholar] [CrossRef]
  21. Ma, Y.; Promthaveepong, K.; Li, N. Chemical sensing on a single SERS particle. ACS Sens. 2017, 2, 135–139. [Google Scholar] [CrossRef]
  22. Huang, C.W.; Hao, Y.W.; Nyagilo, J.; Dave, D.P.; Xu, L.F.; Sun, X.K. Porous hollow gold nanoparticles for cancer SERS imaging. In Journal of Nano Research; Trans Tech Publications Ltd.: Bäch SZ, Switzerland, 2010; pp. 137–148. [Google Scholar]
  23. Pinkhasova, P.; Puccio, B.; Chou, T.; Sukhishvili, S.; Du, H. Noble metal nanostructure both as a SERS nanotag and an analyte probe. Chem. Commun. 2012, 48, 9750–9752. [Google Scholar] [CrossRef]
  24. Emami-Meibodi, M.; Parsaeian, M.; Amraei, R.; Banaei, M.; Anvari, F.; Tahami, S.; Vakhshoor, B.; Mehdizadeh, A.; Nejad, N.F.; Shirmardi, S. An experimental investigation of wastewater treatment using electron beam irradiation. Radiat. Phys. Chem. 2016, 125, 82–87. [Google Scholar] [CrossRef]
  25. Amani, A.M.; Hashemi, S.A.; Mousavi, S.M.; Abrishamifar, S.M.; Vojood, A. Electric field induced alignment of carbon nanotubes: Methodology and outcomes. In Carbon Nanotubes-Recent Progress; IntechOpen: London, UK, 2017. [Google Scholar]
  26. Mousavi, S.M.; Hashemi, S.A.; Jahandideh, S.; Baseri, S.; Zarei, M.; Azadi, S. Modification of phenol novolac epoxy resin and unsaturated polyester using sasobit and silica nanoparticles. Polym. Renew. Resour. 2017, 8, 117–132. [Google Scholar] [CrossRef]
  27. Xu, W.; Ling, X.; Xiao, J.; Dresselhaus, M.S.; Kong, J.; Xu, H.; Liu, Z.; Zhang, J. Surface enhanced Raman spectroscopy on a flat graphene surface. Proc. Natl. Acad. Sci. USA 2012, 109, 9281–9286. [Google Scholar] [CrossRef] [Green Version]
  28. Mousavi, S.M.; Zarei, M.; Hashemi, S.A.; Ramakrishna, S.; Chiang, W.-H.; Lai, C.W.; Gholami, A.; Omidifar, N.; Shokripour, M. Asymmetric membranes: A potential scaffold for wound healing applications. Symmetry 2020, 12, 1100. [Google Scholar] [CrossRef]
  29. Cheng, H.; Zhao, Y.; Fan, Y.; Xie, X.; Qu, L.; Shi, G. Graphene-quantum-dot assembled nanotubes: A new platform for efficient Raman enhancement. Acs Nano 2012, 6, 2237–2244. [Google Scholar] [CrossRef]
  30. Mehrabani, J.; Mousavi, S.; Noaparast, M. Evaluation of the replacement of NaCN with Acidithiobacillus ferrooxidans in the flotation of high-pyrite, low-grade lead–zinc ore. Sep. Purif. Technol. 2011, 80, 202–208. [Google Scholar] [CrossRef]
  31. Hashemi, S.A.; Mousavi, S.M. Effect of bubble based degradation on the physical properties of Single Wall Carbon Nanotube/Epoxy Resin composite and new approach in bubbles reduction. Compos. Part A Appl. Sci. Manuf. 2016, 90, 457–469. [Google Scholar] [CrossRef]
  32. Zheng, X.T.; Ananthanarayanan, A.; Luo, K.Q.; Chen, P. Glowing graphene quantum dots and carbon dots: Properties, syntheses, and biological applications. Small 2015, 11, 1620–1636. [Google Scholar] [CrossRef]
  33. Xu, X.; Ray, R.; Gu, Y.; Ploehn, H.J.; Gearheart, L.; Raker, K.; Scrivens, W.A. Electrophoretic analysis and purification of fluorescent single-walled carbon nanotube fragments. J. Am. Chem. Soc. 2004, 126, 12736–12737. [Google Scholar] [CrossRef]
  34. Ju, S.-Y.; Kopcha, W.P.; Papadimitrakopoulos, F. Brightly fluorescent single-walled carbon nanotubes via an oxygen-excluding surfactant organization. Science 2009, 323, 1319–1323. [Google Scholar] [CrossRef]
  35. Guo, X.; Wang, C.-F.; Yu, Z.-Y.; Chen, L.; Chen, S. Facile access to versatile fluorescent carbon dots toward light-emitting diodes. Chem. Commun. 2012, 48, 2692–2694. [Google Scholar] [CrossRef]
  36. Mousavi, S.M.; Hashemi, S.A.; Zarei, M.; Bahrani, S.; Savardashtaki, A.; Esmaeili, H.; Lai, C.W.; Mazraedoost, S.; Abassi, M.; Ramavandi, B. Data on cytotoxic and antibacterial activity of synthesized Fe3O4 nanoparticles using Malva sylvestris. Data Brief 2020, 28, 104929. [Google Scholar] [CrossRef]
  37. Michalet, X.; Pinaud, F.F.; Bentolila, L.A.; Tsay, J.M.; Doose, S.; Li, J.J.; Sundaresan, G.; Wu, A.; Gambhir, S.; Weiss, S. Quantum dots for live cells, in vivo imaging, and diagnostics. Science 2005, 307, 538–544. [Google Scholar] [CrossRef] [Green Version]
  38. Mousavi, S.; Esmaeili, H.; Arjmand, O.; Karimi, S.; Hashemi, S. Biodegradation study of nanocomposites of phenol novolac epoxy/unsaturated polyester resin/egg shell nanoparticles using natural polymers. J. Mater. 2015, 2015, 131957. [Google Scholar] [CrossRef] [Green Version]
  39. Yu, S.-J.; Kang, M.-W.; Chang, H.-C.; Chen, K.-M.; Yu, Y.-C. Bright fluorescent nanodiamonds: No photobleaching and low cytotoxicity. J. Am. Chem. Soc. 2005, 127, 17604–17605. [Google Scholar] [CrossRef]
  40. Baker, S.N.; Baker, G.A. Luminescent carbon nanodots: Emergent nanolights. Angew. Chem. Int. Ed. 2010, 49, 6726–6744. [Google Scholar] [CrossRef]
  41. Loh, K.P.; Bao, Q.; Eda, G.; Chhowalla, M. Graphene oxide as a chemically tunable platform for optical applications. Nat. Chem. 2010, 2, 1015–1024. [Google Scholar] [CrossRef]
  42. Mousavi, S.; Aghili, A.; Hashemi, S.; Goudarzian, N.; Bakhoda, Z.; Baseri, S. Improved morphology and properties of nanocomposites, linear low density polyethylene, ethylene-co-vinyl acetate and nano clay particles by electron beam. Polym. Renew. Resour. 2016, 7, 135–153. [Google Scholar] [CrossRef]
  43. Wang, C.-F.; Sun, X.-Y.; Su, M.; Wang, Y.-P.; Lv, Y.-K. Electrochemical biosensors based on antibody, nucleic acid and enzyme functionalized graphene for the detection of disease-related biomolecules. Analyst 2020, 145, 1550–1562. [Google Scholar] [CrossRef]
  44. Lin, L.; Rong, M.; Luo, F.; Chen, D.; Wang, Y.; Chen, X. Luminescent graphene quantum dots as new fluorescent materials for environmental and biological applications. TrAC Trends Anal. Chem. 2014, 54, 83–102. [Google Scholar] [CrossRef]
  45. Peng, J.; Gao, W.; Gupta, B.K.; Liu, Z.; Romero-Aburto, R.; Ge, L.; Song, L.; Alemany, L.B.; Zhan, X.; Gao, G. Graphene quantum dots derived from carbon fibers. Nano Lett. 2012, 12, 844–849. [Google Scholar] [CrossRef]
  46. Lu, J.; Yeo, P.S.E.; Gan, C.K.; Wu, P.; Loh, K.P. Transforming C60 molecules into graphene quantum dots. Nat. Nanotechnol. 2011, 6, 247–252. [Google Scholar] [CrossRef]
  47. Sreeprasad, T.; Rodriguez, A.A.; Colston, J.; Graham, A.; Shishkin, E.; Pallem, V.; Berry, V. Electron-tunneling modulation in percolating network of graphene quantum dots: Fabrication, phenomenological understanding, and humidity/pressure sensing applications. Nano Lett. 2013, 13, 1757–1763. [Google Scholar] [CrossRef]
  48. Ye, R.; Xiang, C.; Lin, J.; Peng, Z.; Huang, K.; Yan, Z.; Cook, N.P.; Samuel, E.L.; Hwang, C.-C.; Ruan, G. Coal as an abundant source of graphene quantum dots. Nat. Commun. 2013, 4, 2943. [Google Scholar] [CrossRef]
  49. Yan, Y.; Chen, J.; Li, N.; Tian, J.; Li, K.; Jiang, J.; Liu, J.; Tian, Q.; Chen, P. Systematic bandgap engineering of graphene quantum dots and applications for photocatalytic water splitting and CO2 reduction. ACS Nano 2018, 12, 3523–3532. [Google Scholar] [CrossRef]
  50. Shin, Y.; Park, J.; Hyun, D.; Yang, J.; Lee, J.-H.; Kim, J.-H.; Lee, H. Acid-free and oxone oxidant-assisted solvothermal synthesis of graphene quantum dots using various natural carbon materials as resources. Nanoscale 2015, 7, 5633–5637. [Google Scholar] [CrossRef]
  51. Zhang, Y.; Li, K.; Ren, S.; Dang, Y.; Liu, G.; Zhang, R.; Zhang, K.; Long, X.; Jia, K. Coal-derived graphene quantum dots produced by ultrasonic physical tailoring and their capacity for Cu (II) detection. ACS Sustain. Chem. Eng. 2019, 7, 9793–9799. [Google Scholar] [CrossRef]
  52. Bacon, M.; Bradley, S.J.; Nann, T. Graphene quantum dots. Part. Part. Syst. Charact. 2014, 31, 415–428. [Google Scholar] [CrossRef]
  53. Shen, J.; Zhu, Y.; Chen, C.; Yang, X.; Li, C. Facile preparation and upconversion luminescence of graphene quantum dots. Chem. Commun. 2011, 47, 2580–2582. [Google Scholar] [CrossRef]
  54. Li, Y.; Hu, Y.; Zhao, Y.; Shi, G.; Deng, L.; Hou, Y.; Qu, L. An electrochemical avenue to green-luminescent graphene quantum dots as potential electron-acceptors for photovoltaics. Adv. Mater. 2011, 23, 776–780. [Google Scholar] [CrossRef]
  55. Lee, E.; Ryu, J.; Jang, J. Fabrication of graphene quantum dots via size-selective precipitation and their application in upconversion-based DSSCs. Chem. Commun. 2013, 49, 9995–9997. [Google Scholar] [CrossRef]
  56. Dong, Y.; Shao, J.; Chen, C.; Li, H.; Wang, R.; Chi, Y.; Lin, X.; Chen, G. Blue luminescent graphene quantum dots and graphene oxide prepared by tuning the carbonization degree of citric acid. Carbon 2012, 50, 4738–4743. [Google Scholar] [CrossRef]
  57. Wu, Z.L.; Gao, M.X.; Wang, T.T.; Wan, X.Y.; Zheng, L.L.; Huang, C.Z. A general quantitative pH sensor developed with dicyandiamide N-doped high quantum yield graphene quantum dots. Nanoscale 2014, 6, 3868–3874. [Google Scholar] [CrossRef]
  58. Kumar, S.; Aziz, S.T.; Girshevitz, O.; Nessim, G.D. One-step synthesis of N-doped graphene quantum dots from chitosan as a sole precursor using chemical vapor deposition. J. Phys. Chem. C 2018, 122, 2343–2349. [Google Scholar] [CrossRef]
  59. Kurniawan, D.; Chiang, W.-H. Microplasma-enabled colloidal nitrogen-doped graphene quantum dots for broad-range fluorescent pH sensors. Carbon 2020, 167, 675–684. [Google Scholar] [CrossRef]
  60. Kurniawan, D.; Anjali, B.A.; Setiawan, O.; Ostrikov, K.K.; Chung, Y.G.; Chiang, W.-H. Microplasma Band Structure Engineering in Graphene Quantum Dots for Sensitive and Wide-Range pH Sensing. ACS Appl. Mater. Interfaces 2021, 14, 1670–1683. [Google Scholar] [CrossRef]
  61. Ding, Z.; Li, F.; Wen, J.; Wang, X.; Sun, R. Gram-scale synthesis of single-crystalline graphene quantum dots derived from lignin biomass. Green Chem. 2018, 20, 1383–1390. [Google Scholar] [CrossRef]
  62. Kurniawan, D.; Weng, R.-J.; Setiawan, O.; Ostrikov, K.K.; Chiang, W.-H. Microplasma nanoengineering of emission-tuneable colloidal nitrogen-doped graphene quantum dots as smart environmental-responsive nanosensors and nanothermometers. Carbon 2021, 185, 501–513. [Google Scholar] [CrossRef]
  63. Dong, Y.; Chen, C.; Zheng, X.; Gao, L.; Cui, Z.; Yang, H.; Guo, C.; Chi, Y.; Li, C.M. One-step and high yield simultaneous preparation of single-and multi-layer graphene quantum dots from CX-72 carbon black. J. Mater. Chem. 2012, 22, 8764–8766. [Google Scholar] [CrossRef]
  64. Pan, D.; Guo, L.; Zhang, J.; Xi, C.; Xue, Q.; Huang, H.; Li, J.; Zhang, Z.; Yu, W.; Chen, Z. Cutting sp 2 clusters in graphene sheets into colloidal graphene quantum dots with strong green fluorescence. J. Mater. Chem. 2012, 22, 3314–3318. [Google Scholar] [CrossRef]
  65. Li, L.L.; Ji, J.; Fei, R.; Wang, C.Z.; Lu, Q.; Zhang, J.R.; Jiang, L.P.; Zhu, J.J. A facile microwave avenue to electrochemiluminescent two-color graphene quantum dots. Adv. Funct. Mater. 2012, 22, 2971–2979. [Google Scholar] [CrossRef]
  66. Chen, S.; Liu, J.-W.; Chen, M.-L.; Chen, X.-W.; Wang, J.-H. Unusual emission transformation of graphene quantum dots induced by self-assembled aggregation. Chem. Commun. 2012, 48, 7637–7639. [Google Scholar] [CrossRef]
  67. Shinde, D.B.; Pillai, V.K. Electrochemical preparation of luminescent graphene quantum dots from multiwalled carbon nanotubes. Chem. –A Eur. J. 2012, 18, 12522–12528. [Google Scholar] [CrossRef]
  68. Tang, L.; Ji, R.; Cao, X.; Lin, J.; Jiang, H.; Li, X.; Teng, K.S.; Luk, C.M.; Zeng, S.; Hao, J. Deep ultraviolet photoluminescence of water-soluble self-passivated graphene quantum dots. ACS Nano 2012, 6, 5102–5110. [Google Scholar] [CrossRef]
  69. Liu, R.; Wu, D.; Feng, X.; Müllen, K. Bottom-up fabrication of photoluminescent graphene quantum dots with uniform morphology. J. Am. Chem. Soc. 2011, 133, 15221–15223. [Google Scholar] [CrossRef]
  70. Mangolini, L.; Thimsen, E.; Kortshagen, U. High-yield plasma synthesis of luminescent silicon nanocrystals. Nano Lett. 2005, 5, 655–659. [Google Scholar] [CrossRef]
  71. Knipping, J.; Wiggers, H.; Rellinghaus, B.; Roth, P.; Konjhodzic, D.; Meier, C. Synthesis of high purity silicon nanoparticles in a low pressure microwave reactor. J. Nanosci. Nanotechnol. 2004, 4, 1039–1044. [Google Scholar] [CrossRef]
  72. Sankaran, R.M.; Holunga, D.; Flagan, R.C.; Giapis, K.P. Synthesis of blue luminescent Si nanoparticles using atmospheric-pressure microdischarges. Nano Lett. 2005, 5, 537–541. [Google Scholar] [CrossRef] [Green Version]
  73. Kortshagen, U. Nonthermal plasma synthesis of semiconductor nanocrystals. J. Phys. D Appl. Phys. 2009, 42, 113001. [Google Scholar] [CrossRef]
  74. Pi, X.; Kortshagen, U. Nonthermal plasma synthesized freestanding silicon–germanium alloy nanocrystals. Nanotechnology 2009, 20, 295602. [Google Scholar] [CrossRef]
  75. Pi, X.; Gresback, R.; Liptak, R.; Campbell, S.; Kortshagen, U. Doping efficiency, dopant location, and oxidation of Si nanocrystals. Appl. Phys. Lett. 2008, 92, 123102. [Google Scholar] [CrossRef]
  76. Ni, Z.; Pi, X.; Ali, M.; Zhou, S.; Nozaki, T.; Yang, D. Freestanding doped silicon nanocrystals synthesized by plasma. J. Phys. D Appl. Phys. 2015, 48, 314006. [Google Scholar] [CrossRef]
  77. Pereira, R.; Almeida, A. Doped semiconductor nanoparticles synthesized in gas-phase plasmas. J. Phys. D Appl. Phys. 2015, 48, 314005. [Google Scholar] [CrossRef]
  78. Pi, X.; Yu, T.; Yang, D. Water-Dispersible Silicon-Quantum-Dot-Containing Micelles Self-Assembled from an Amphiphilic Polymer. Part. Part. Syst. Charact. 2014, 31, 751–756. [Google Scholar] [CrossRef]
  79. Mangolini, L.; Kortshagen, U. Plasma-assisted synthesis of silicon nanocrystal inks. Adv. Mater. 2007, 19, 2513–2519. [Google Scholar] [CrossRef]
  80. Mousavi, S.M.; Hashemi, S.A.; Kalashgrani, M.Y.; Omidifar, N.; Bahrani, S.; Vijayakameswara Rao, N.; Babapoor, A.; Gholami, A.; Chiang, W.-H. Bioactive Graphene Quantum Dots Based Polymer Composite for Biomedical Applications. Polymers 2022, 14, 617. [Google Scholar] [CrossRef] [PubMed]
  81. Pignatello, R. Biomaterials Science and Engineering; BoD–Books on Demand: Norderstedt, Germany, 2011. [Google Scholar]
  82. Horst, F.H.; da Silva Rodrigues, C.V.; Carvalho, P.H.P.R.; Leite, A.M.; Azevedo, R.B.; Neto, B.A.; Corrêa, J.R.; Garcia, M.P.; Alotaibi, S.; Henini, M. From cow manure to bioactive carbon dots: A light-up probe for bioimaging investigations, glucose detection and potential immunotherapy agent for melanoma skin cancer. RSC Adv. 2021, 11, 6346–6352. [Google Scholar] [CrossRef] [PubMed]
  83. Mousavi, S.M.; Hashemi, S.A.; Ramakrishna, S.; Esmaeili, H.; Bahrani, S.; Koosha, M.; Babapoor, A. Green synthesis of supermagnetic Fe3O4–MgO nanoparticles via Nutmeg essential oil toward superior anti-bacterial and anti-fungal performance. J. Drug Deliv. Sci. Technol. 2019, 54, 101352. [Google Scholar] [CrossRef]
  84. Han, Y.; Zhang, F.; Zhang, J.; Shao, D.; Wang, Y.; Li, S.; Lv, S.; Chi, G.; Zhang, M.; Chen, L. Bioactive carbon dots direct the osteogenic differentiation of human bone marrow mesenchymal stem cells. Colloids Surf. B Biointerfaces 2019, 179, 1–8. [Google Scholar] [CrossRef]
  85. Seyed, M.M. Unsaturated polyester resins modified with cresol novolac epoxy and silica nanoparticles: Processing and mechanical properties. Int. J. Chem. Pet. Sci. (IJCPS) 2016, 5, 13–26. [Google Scholar]
  86. Abbas, A.; Mariana, L.T.; Phan, A.N. Biomass-waste derived graphene quantum dots and their applications. Carbon 2018, 140, 77–99. [Google Scholar] [CrossRef] [Green Version]
  87. Kalita, H.; Mohapatra, J.; Pradhan, L.; Mitra, A.; Bahadur, D.; Aslam, M. Efficient synthesis of rice based graphene quantum dots and their fluorescent properties. RSC Adv. 2016, 6, 23518–23524. [Google Scholar] [CrossRef]
  88. Roy, P.; Periasamy, A.P.; Chuang, C.; Liou, Y.-R.; Chen, Y.-F.; Joly, J.; Liang, C.-T.; Chang, H.-T. Plant leaf-derived graphene quantum dots and applications for white LEDs. New J. Chem. 2014, 38, 4946–4951. [Google Scholar] [CrossRef]
  89. Nirala, N.R.; Khandelwal, G.; Kumar, B.; Prakash, R.; Kumar, V. One step electro-oxidative preparation of graphene quantum dots from wood charcoal as a peroxidase mimetic. Talanta 2017, 173, 36–43. [Google Scholar] [CrossRef]
  90. Suryawanshi, A.; Biswal, M.; Mhamane, D.; Gokhale, R.; Patil, S.; Guin, D.; Ogale, S. Large scale synthesis of graphene quantum dots (GQDs) from waste biomass and their use as an efficient and selective photoluminescence on–off–on probe for Ag+ ions. Nanoscale 2014, 6, 11664–11670. [Google Scholar] [CrossRef]
  91. Wang, L.; Li, W.; Wu, B.; Li, Z.; Wang, S.; Liu, Y.; Pan, D.; Wu, M. Facile synthesis of fluorescent graphene quantum dots from coffee grounds for bioimaging and sensing. Chem. Eng. J. 2016, 300, 75–82. [Google Scholar] [CrossRef]
  92. Lagarón, J.-M. Multifunctional and nanoreinforced polymers for food packaging. In Multifunctional and Nanoreinforced Polymers for Food Packaging; Elsevier: Amsterdam, The Netherlands, 2011; pp. 1–28. [Google Scholar]
  93. Wang, J.; Zhang, H.; Wang, F.; Ai, X.; Huang, D.; Liu, G.; Mi, P. Enzyme-responsive polymers for drug delivery and molecular imaging. In Stimuli Responsive Polymeric Nanocarriers for Drug Delivery Applications, Volume 1; Elsevier: Amsterdam, The Netherlands, 2018; pp. 101–119. [Google Scholar]
  94. Hung, L.X.; Bassène, P.D.; Thang, P.N.; Loan, N.T.; de Marcillac, W.D.; Dhawan, A.R.; Feng, F.; Esparza-Villa, J.U.; Hien, N.T.T.; Liem, N.Q. Near-infrared emitting CdTeSe alloyed quantum dots: Raman scattering, photoluminescence and single-emitter optical properties. RSC Adv. 2017, 7, 47966–47974. [Google Scholar] [CrossRef] [Green Version]
  95. Valappil, M.O.; Pillai, V.K.; Alwarappan, S. Spotlighting graphene quantum dots and beyond: Synthesis, properties and sensing applications. Appl. Mater. Today 2017, 9, 350–371. [Google Scholar] [CrossRef]
  96. Kleinman, S.L.; Frontiera, R.R.; Henry, A.-I.; Dieringer, J.A.; Van Duyne, R.P. Creating, characterizing, and controlling chemistry with SERS hot spots. Phys. Chem. Chem. Phys. 2013, 15, 21–36. [Google Scholar] [CrossRef]
  97. Nie, S.; Emory, S.R. Probing single molecules and single nanoparticles by surface-enhanced Raman scattering. Science 1997, 275, 1102–1106. [Google Scholar] [CrossRef]
  98. Kneipp, K.; Wang, Y.; Kneipp, H.; Perelman, L.T.; Itzkan, I.; Dasari, R.R.; Feld, M.S. Single molecule detection using surface-enhanced Raman scattering (SERS). Phys. Rev. Lett. 1997, 78, 1667. [Google Scholar] [CrossRef] [Green Version]
  99. Zhang, W.; Li, C.; Gao, K.; Lu, F.; Liu, M.; Li, X.; Zhang, L.; Mao, D.; Gao, F.; Huang, L. Surface-enhanced Raman spectroscopy with Au-nanoparticle substrate fabricated by using femtosecond pulse. Nanotechnology 2018, 29, 205301. [Google Scholar] [CrossRef]
  100. Chang, T.-H.; Chang, Y.-C.; Chen, C.-M.; Chuang, K.-W.; Chou, C.-M. A facile method to directly deposit the large-scale Ag nanoparticles on a silicon substrate for sensitive, uniform, reproducible and stable SERS substrate. J. Alloys Compd. 2019, 782, 887–892. [Google Scholar] [CrossRef]
  101. Wu, H.; Luo, Y.; Hou, C.; Huo, D.; Zhou, Y.; Zou, S.; Zhao, J.; Lei, Y. Flexible bipyramid-AuNPs based SERS tape sensing strategy for detecting methyl parathion on vegetable and fruit surface. Sens. Actuators B Chem. 2019, 285, 123–128. [Google Scholar] [CrossRef]
  102. Fazio, E.; Neri, F.; Savasta, S.; Spadaro, S.; Trusso, S. Surface-enhanced Raman scattering of SnO2 bulk material and colloidal solutions. Phys. Rev. B 2012, 85, 195423. [Google Scholar] [CrossRef] [Green Version]
  103. Doğan, İ.; Gresback, R.; Nozaki, T.; van de Sanden, M. Analysis of temporal evolution of quantum dot surface chemistry by surface-enhanced Raman scattering. Sci. Rep. 2016, 6, 29508. [Google Scholar] [CrossRef] [Green Version]
  104. Milekhin, A.G.; Sveshnikova, L.; Duda, T.; Surovtsev, N.V.; Adichtchev, S.; Zahn, D.R. Surface enhanced Raman scattering by CdS quantum dots. JETP Lett. 2008, 88, 799–801. [Google Scholar] [CrossRef]
  105. Stadelmann, K.; Elizabeth, A.; Sabanés, N.M.; Domke, K.F. The SERS signature of PbS quantum dot oxidation. Vib. Spectrosc. 2017, 91, 157–162. [Google Scholar] [CrossRef] [Green Version]
  106. Hugall, J.T.; Baumberg, J.J.; Mahajan, S. Surface-enhanced Raman spectroscopy of CdSe quantum dots on nanostructured plasmonic surfaces. Appl. Phys. Lett. 2009, 95, 141111. [Google Scholar] [CrossRef] [Green Version]
  107. Lee, Y.-b.; Ho Lee, S.; Lee, S.; Lee, H.; Kim, J.; Joo, J. Surface enhanced Raman scattering effect of CdSe/ZnS quantum dots hybridized with Au nanowire. Appl. Phys. Lett. 2013, 102, 033109. [Google Scholar] [CrossRef]
  108. Sheremet, E.; Milekhin, A.; Rodriguez, R.; Weiss, T.; Nesterov, M.; Rodyakina, E.; Gordan, O.; Sveshnikova, L.; Duda, T.; Gridchin, V. Surface-and tip-enhanced resonant Raman scattering from CdSe nanocrystals. Phys. Chem. Chem. Phys. 2015, 17, 21198–21203. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Bak, S.; Kim, D.; Lee, H. Graphene quantum dots and their possible energy applications: A review. Curr. Appl. Phys. 2016, 16, 1192–1201. [Google Scholar] [CrossRef]
  110. Panyathip, R.; Sucharitakul, S.; Phaduangdhitidhada, S.; Ngamjarurojana, A.; Kumnorkaew, P.; Choopun, S. Surface Enhanced Raman Scattering in Graphene Quantum Dots Grown via Electrochemical Process. Molecules 2021, 26, 5484. [Google Scholar] [CrossRef] [PubMed]
  111. Ostrovskaya, L.Y. Characterization of different carbon nanomaterials promising for biomedical and sensor applications by the wetting method. Powder Metall. Met. Ceram. 2003, 42, 1–8. [Google Scholar] [CrossRef]
  112. Ling, X.; Xie, L.; Fang, Y.; Xu, H.; Zhang, H.; Kong, J.; Dresselhaus, M.S.; Zhang, J.; Liu, Z. Can graphene be used as a substrate for Raman enhancement? Nano Lett. 2010, 10, 553–561. [Google Scholar] [CrossRef] [PubMed]
  113. Gulzar, A.; Yang, P.; He, F.; Xu, J.; Yang, D.; Xu, L.; Jan, M.O. Bioapplications of graphene constructed functional nanomaterials. Chem. -Biol. Interact. 2017, 262, 69–89. [Google Scholar] [CrossRef]
  114. Kumar, S.; Tokunaga, K.; Namura, K.; Fukuoka, T.; Suzuki, M. Experimental evidence of a twofold electromagnetic enhancement mechanism of surface-enhanced Raman scattering. J. Phys. Chem. C 2020, 124, 21215–21222. [Google Scholar] [CrossRef]
  115. Lu, Y.; Lin, L.; Ye, J. Human metabolite detection by surface-enhanced Raman spectroscopy. Mater. Today Bio 2022, 13, 100205. [Google Scholar] [CrossRef]
  116. Pearson, B.; Wang, P.; Mills, A.; Pang, S.; McLandsborough, L.; He, L. Innovative sandwich assay with dual optical and SERS sensing mechanisms for bacterial detection. Anal. Methods 2017, 9, 4732–4739. [Google Scholar] [CrossRef]
  117. Fateixa, S.; Nogueira, H.I.; Trindade, T. Hybrid nanostructures for SERS: Materials development and chemical detection. Phys. Chem. Chem. Phys. 2015, 17, 21046–21071. [Google Scholar] [CrossRef]
  118. Sharma, B.; Frontiera, R.R.; Henry, A.-I.; Ringe, E.; Van Duyne, R.P. SERS: Materials, applications, and the future. Mater. Today 2012, 15, 16–25. [Google Scholar] [CrossRef]
  119. Ngo, H.T.; Wang, H.-N.; Fales, A.M.; Vo-Dinh, T. Plasmonic SERS biosensing nanochips for DNA detection. Anal. Bioanal. Chem. 2016, 408, 1773–1781. [Google Scholar] [CrossRef] [PubMed]
  120. Ganesh, S.; Venkatakrishnan, K.; Tan, B. Quantum scale organic semiconductors for SERS detection of DNA methylation and gene expression. Nat. Commun. 2020, 11, 1135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  121. Saviñon-Flores, F.; Méndez, E.; López-Castaños, M.; Carabarin-Lima, A.; López-Castaños, K.A.; González-Fuentes, M.A.; Méndez-Albores, A. A review on SERS-based detection of human virus infections: Influenza and coronavirus. Biosensors 2021, 11, 66. [Google Scholar] [CrossRef]
  122. Rajender, G.; Giri, P. Formation mechanism of graphene quantum dots and their edge state conversion probed by photoluminescence and Raman spectroscopy. J. Mater. Chem. C 2016, 4, 10852–10865. [Google Scholar] [CrossRef]
  123. Aronson, J.K. Biomarkers and surrogate endpoints. Br. J. Clin. Pharmacol. 2005, 59, 491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  124. Tanaka, T.; Tanaka, M.; Tanaka, T.; Ishigamori, R. Biomarkers for colorectal cancer. Int. J. Mol. Sci. 2010, 11, 3209–3225. [Google Scholar] [CrossRef] [Green Version]
  125. Vaughan, L. Biomarkers in acute medicine. Medicine 2017, 45, 150–156. [Google Scholar] [CrossRef]
  126. Mousavi, S.M.; Low, F.W.; Hashemi, S.A.; Lai, C.W.; Ghasemi, Y.; Soroshnia, S.; Savardashtaki, A.; Babapoor, A.; Pynadathu Rumjit, N.; Goh, S.M. Development of graphene based nanocomposites towards medical and biological applications. Artif. Cells Nanomed. Biotechnol. 2020, 48, 1189–1205. [Google Scholar] [CrossRef]
  127. McCormick, T.; Martin, K.; Hehenberger, M. The Evolving Role of Biomarkers: Focusing on Patients from Research to Clinical Practice; IBM Global Business Services: Pyrmont, NSW, Australia, 2007; Volume 4, pp. 1–20. [Google Scholar]
  128. Mayeux, R. Biomarkers: Potential uses and limitations. NeuroRx 2004, 1, 182–188. [Google Scholar] [CrossRef]
  129. Gârban, Z.; Avacovici, A.; Gârban, G.; Ghibu, G.; Velciov, A.B.; Pop, C.I. Biomarkers: Theoretical aspects and applicative peculiarities note i. general characteristics of biomarkers. Agroaliment. Process Technol 2005, 11, 139–146. [Google Scholar]
  130. Hashemi, S.A.; Mousavi, S.M.; Faghihi, R.; Arjmand, M.; Rahsepar, M.; Bahrani, S.; Ramakrishna, S.; Lai, C.W. Superior X-ray radiation shielding effectiveness of biocompatible polyaniline reinforced with hybrid graphene oxide-iron tungsten nitride flakes. Polymers 2020, 12, 1407. [Google Scholar] [CrossRef] [PubMed]
  131. Polivka, J.; Krakorova, K.; Peterka, M.; Topolcan, O. Current status of biomarker research in neurology. EPMA J. 2016, 7, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  132. Tietjen, G.E.; Khubchandani, J. Vascular biomarkers in migraine. Cephalalgia 2015, 35, 95–117. [Google Scholar] [CrossRef]
  133. Tietjen, G.E.; Khubchandani, J.; Herial, N.A.; Shah, K. Adverse childhood experiences are associated with migraine and vascular biomarkers. Headache: J. Head Face Pain 2012, 52, 920–929. [Google Scholar] [CrossRef]
  134. Durham, P.; Papapetropoulos, S. Biomarkers associated with migraine and their potential role in migraine management. Headache J. Head Face Pain 2013, 53, 1262–1277. [Google Scholar] [CrossRef]
  135. Hashemi, S.A.; Mousavi, S.M.; Naderi, H.R.; Bahrani, S.; Arjmand, M.; Hagfeldt, A.; Chiang, W.-H.; Ramakrishna, S. Reinforced polypyrrole with 2D graphene flakes decorated with interconnected nickel-tungsten metal oxide complex toward superiorly stable supercapacitor. Chem. Eng. J. 2021, 418, 129396. [Google Scholar] [CrossRef]
  136. Ishizaki, K.; Takeshima, T.; Fukuhara, Y.; Araki, H.; Nakaso, K.; Kusumi, M.; Nakashima, K. Increased plasma transforming growth factor-β1 in migraine. Headache J. Head Face Pain 2005, 45, 1224–1228. [Google Scholar] [CrossRef]
  137. Kurth, T.; Ridker, P.; Buring, J. Migraine and biomarkers of cardiovascular disease in women. Cephalalgia 2008, 28, 49–56. [Google Scholar] [CrossRef]
  138. Song, C.; Guo, S.; Jin, S.; Chen, L.; Jung, Y.M. Biomarkers determination based on surface-enhanced Raman scattering. Chemosensors 2020, 8, 118. [Google Scholar] [CrossRef]
  139. Li, Y.; Liu, X.; Guo, J.; Zhang, Y.; Guo, J.; Wu, X.; Wang, B.; Ma, X. Simultaneous detection of inflammatory biomarkers by SERS nanotag-based lateral flow assay with portable cloud Raman spectrometer. Nanomaterials 2021, 11, 1496. [Google Scholar] [CrossRef] [PubMed]
  140. Ranjan, R.; Esimbekova, E.N.; Kratasyuk, V.A. Rapid biosensing tools for cancer biomarkers. Biosens. Bioelectron. 2017, 87, 918–930. [Google Scholar] [CrossRef]
  141. Chen, R.; Du, X.; Cui, Y.; Zhang, X.; Ge, Q.; Dong, J.; Zhao, X. Vertical flow assay for inflammatory biomarkers based on nanofluidic channel array and SERS nanotags. Small 2020, 16, 2002801. [Google Scholar] [CrossRef]
  142. Liu, H.; Gao, X.; Xu, C.; Liu, D. SERS Tags for Biomedical Detection and Bioimaging. Theranostics 2022, 12, 1870. [Google Scholar] [CrossRef] [PubMed]
  143. Singh, S.; Deshmukh, A.; Chaturvedi, P.; Krishna, C.M. In vivo Raman spectroscopic identification of premalignant lesions in oral buccal mucosa. J. Biomed. Opt. 2012, 17, 105002. [Google Scholar] [CrossRef] [PubMed]
  144. Hamza, T.H.; Zabetian, C.P.; Tenesa, A.; Laederach, A.; Montimurro, J.; Yearout, D.; Kay, D.M.; Doheny, K.F.; Paschall, J.; Pugh, E. Common genetic variation in the HLA region is associated with late-onset sporadic Parkinson’s disease. Nat. Genet. 2010, 42, 781–785. [Google Scholar] [CrossRef]
  145. Araujo, D.; Lapchak, P. Induction of immune system mediators in the hippocampal formation in Alzheimer’s and Parkinson’s diseases: Selective effects on specific interleukins and interleukin receptors. Neuroscience 1994, 61, 745–754. [Google Scholar] [CrossRef]
  146. Maddon, P.J.; Dalgleish, A.G.; McDougal, J.S.; Clapham, P.R.; Weiss, R.A.; Axel, R. The T4 gene encodes the AIDS virus receptor and is expressed in the immune system and the brain. Cell 1986, 47, 333–348. [Google Scholar] [CrossRef]
  147. Maffia, P.; Zinselmeyer, B.H.; Ialenti, A.; Kennedy, S.; Baker, A.H.; McInnes, I.B.; Brewer, J.M.; Garside, P. Images in cardiovascular medicine: Multiphoton microscopy for three-dimensional imaging of lymphocyte recruitment into apolipoprotein-E-deficient mouse carotid artery. Circulation 2007, 115, e326–e328. [Google Scholar] [CrossRef] [Green Version]
  148. Gabriel, S.E. The epidemiology of rheumatoid arthritis. Rheum. Dis. Clin. N. Am. 2001, 27, 269–281. [Google Scholar] [CrossRef]
  149. Ortolano, F.; Maffia, P.; Dever, G.; Hutchison, S.; Benson, R.; Millington, O.; De Simoni, M.; Bushell, T.; Garside, P.; Carswell, H. Imaging T-cell movement in the brain during experimental cerebral malaria. Parasite Immunol. 2009, 31, 147–150. [Google Scholar] [CrossRef]
  150. Fumagalli, S.; Coles, J.A.; Ejlerskov, P.; Ortolano, F.; Bushell, T.J.; Brewer, J.M.; De Simoni, M.-G.; Dever, G.; Garside, P.; Maffia, P. In vivo real-time multiphoton imaging of T lymphocytes in the mouse brain after experimental stroke. Stroke 2011, 42, 1429–1436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  151. Theer, P.; Denk, W. On the fundamental imaging-depth limit in two-photon microscopy. JOSA A 2006, 23, 3139–3149. [Google Scholar] [CrossRef] [PubMed]
  152. Zipfel, W.R.; Williams, R.M.; Christie, R.; Nikitin, A.Y.; Hyman, B.T.; Webb, W.W. Live tissue intrinsic emission microscopy using multiphoton-excited native fluorescence and second harmonic generation. Proc. Natl. Acad. Sci. USA 2003, 100, 7075–7080. [Google Scholar] [CrossRef] [Green Version]
  153. Patterson, G.H.; Piston, D.W. Photobleaching in two-photon excitation microscopy. Biophys. J. 2000, 78, 2159–2162. [Google Scholar] [CrossRef] [Green Version]
  154. König, K. Multiphoton microscopy in life sciences. J. Microsc. 2000, 200, 83–104. [Google Scholar]
  155. McQueenie, R.; Stevenson, R.; Benson, R.; MacRitchie, N.; McInnes, I.; Maffia, P.; Faulds, K.; Graham, D.; Brewer, J.; Garside, P. Detection of inflammation in vivo by surface-enhanced Raman scattering provides higher sensitivity than conventional fluorescence imaging. Anal. Chem. 2012, 84, 5968–5975. [Google Scholar] [CrossRef]
  156. Zavaleta, C.L.; Smith, B.R.; Walton, I.; Doering, W.; Davis, G.; Shojaei, B.; Natan, M.J.; Gambhir, S.S. Multiplexed imaging of surface enhanced Raman scattering nanotags in living mice using noninvasive Raman spectroscopy. Proc. Natl. Acad. Sci. USA 2009, 106, 13511–13516. [Google Scholar]
  157. Maiti, K.K.; Dinish, U.; Fu, C.Y.; Lee, J.-J.; Soh, K.-S.; Yun, S.-W.; Bhuvaneswari, R.; Olivo, M.; Chang, Y.-T. Development of biocompatible SERS nanotag with increased stability by chemisorption of reporter molecule for in vivo cancer detection. Biosens. Bioelectron. 2010, 26, 398–403. [Google Scholar]
Figure 1. Graphene quantum dot. ( Biosensors 12 00461 i001: Carbon, Biosensors 12 00461 i002: Hydrogen, Biosensors 12 00461 i003: Oxygen).
Figure 1. Graphene quantum dot. ( Biosensors 12 00461 i001: Carbon, Biosensors 12 00461 i002: Hydrogen, Biosensors 12 00461 i003: Oxygen).
Biosensors 12 00461 g001
Figure 2. Schematic diagram representing the top-down and bottom-up approaches for the synthesis of GQDs.
Figure 2. Schematic diagram representing the top-down and bottom-up approaches for the synthesis of GQDs.
Biosensors 12 00461 g002
Figure 3. (a) The synthesis of GQD (C57H26O11) from different types of biomass-waste; (b) figure illustrates the various approaches used for converting biowaste into GQDs, along with how they can be used as energy sources.
Figure 3. (a) The synthesis of GQD (C57H26O11) from different types of biomass-waste; (b) figure illustrates the various approaches used for converting biowaste into GQDs, along with how they can be used as energy sources.
Biosensors 12 00461 g003
Figure 4. (a) Mechanism of the surface-enhanced Raman scattering (SERS) electromagnetic (EM) effect, electromagnetic SERS enhancement. Reprinted with permission from [114]. Copyright © 2020, American Chemical Society. (b) Schematic illustration of molecules on graphene and a substrate and Raman experiments. Reprinted with permission from [112], Copyright © 2010, American Chemical Society.
Figure 4. (a) Mechanism of the surface-enhanced Raman scattering (SERS) electromagnetic (EM) effect, electromagnetic SERS enhancement. Reprinted with permission from [114]. Copyright © 2020, American Chemical Society. (b) Schematic illustration of molecules on graphene and a substrate and Raman experiments. Reprinted with permission from [112], Copyright © 2010, American Chemical Society.
Biosensors 12 00461 g004
Figure 5. In vivo imaging of SERS-enhanced GQDs for detection of tumours, inflammation and living tissues.
Figure 5. In vivo imaging of SERS-enhanced GQDs for detection of tumours, inflammation and living tissues.
Biosensors 12 00461 g005
Figure 6. Schematic of C-reactive protein and cytokines the functional pathways of CRP.
Figure 6. Schematic of C-reactive protein and cytokines the functional pathways of CRP.
Biosensors 12 00461 g006
Figure 7. The major processes through which adiponectin maintains metabolic homeostasis.
Figure 7. The major processes through which adiponectin maintains metabolic homeostasis.
Biosensors 12 00461 g007
Figure 8. Detection of inflammatory by SERS-enhanced GQDs.
Figure 8. Detection of inflammatory by SERS-enhanced GQDs.
Biosensors 12 00461 g008
Table 1. Characteristics of top-down and bottom-up methods in the synthesis of GQDs.
Table 1. Characteristics of top-down and bottom-up methods in the synthesis of GQDs.
SubgroupInitial MaterialSize (nm)Quantum EfficiencyRef.
Top-downAcid oxidationCarbon black1544.5[63]
HydrothermalGraphene oxide5–135[64]
SolvothermalGraphene oxide3–51.6[65]
MicrowaveGraphene oxide2–78[66]
Ultrasound wavesGraphene3–5-[54]
ElectrochemicalGraphite5–10-[67]
Bottom-upPyrolysis of the precursorGlucose1.65–21-[68]
Catalytic opening of the cageFullerene 602.7–1015–30[46]
PyrolysisHexa benzo chromen~60-[69]
Table 2. Synthesis of GQDs from different types of biomass-waste.
Table 2. Synthesis of GQDs from different types of biomass-waste.
PrecursorProductPreparation ApproachSize (nm)Ref.
Rice grainsGQDsPyrolysis2–6.5[87]
Fenugreek leaf extractGQDsPyrolysis and hydrothermal treatment3–10[88]
Wood charcoalGQDsElectrochemical oxidation3–6[89]
Neem leavesGQDs, Am-GQDsPyrolysis and hydrothermal treatments5–6[90]
Coffee groundsGQDs, PEIGQDsHydrothermal treatment1.88 (GQDs)2.67 (PEIGQDs)[91]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mousavi, S.M.; Hashemi, S.A.; Yari Kalashgrani, M.; Kurniawan, D.; Gholami, A.; Rahmanian, V.; Omidifar, N.; Chiang, W.-H. Recent Advances in Inflammatory Diagnosis with Graphene Quantum Dots Enhanced SERS Detection. Biosensors 2022, 12, 461. https://doi.org/10.3390/bios12070461

AMA Style

Mousavi SM, Hashemi SA, Yari Kalashgrani M, Kurniawan D, Gholami A, Rahmanian V, Omidifar N, Chiang W-H. Recent Advances in Inflammatory Diagnosis with Graphene Quantum Dots Enhanced SERS Detection. Biosensors. 2022; 12(7):461. https://doi.org/10.3390/bios12070461

Chicago/Turabian Style

Mousavi, Seyyed Mojtaba, Seyyed Alireza Hashemi, Masoomeh Yari Kalashgrani, Darwin Kurniawan, Ahmad Gholami, Vahid Rahmanian, Navid Omidifar, and Wei-Hung Chiang. 2022. "Recent Advances in Inflammatory Diagnosis with Graphene Quantum Dots Enhanced SERS Detection" Biosensors 12, no. 7: 461. https://doi.org/10.3390/bios12070461

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