Next Article in Journal
Carbon Nanohorns as Effective Nanotherapeutics in Cancer Therapy
Next Article in Special Issue
Functionalized Graphene Quantum Dots Modulate Malignancy of Glioblastoma Multiforme by Downregulating Neurospheres Formation
Previous Article in Journal
Effect of Molybdenum Content on Mechanical and Tribological Properties of Diamond-Like Carbon Coatings over Titanium β-21S Alloy
Previous Article in Special Issue
Nitrogen-Doped Arginine Carbon Dots and Its Metal Nanoparticle Composites as Antibacterial Agent
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Citric Acid Derived Carbon Dots, the Challenge of Understanding the Synthesis-Structure Relationship

Laboratory of Materials Science and Nanotechnology (LMNT), Department of Chemistry and Pharmacy, University of Sassari, CR-INSTM, Via Vienna 2, 07100 Sassari, Italy
*
Author to whom correspondence should be addressed.
Submission received: 2 December 2020 / Revised: 16 December 2020 / Accepted: 17 December 2020 / Published: 22 December 2020
(This article belongs to the Special Issue Graphene and Carbon Quantum Dots, and Related 2D Quantum Dots)

Abstract

:
Carbon dots (CDs) are highly-emissive nanoparticles obtained through fast and cheap syntheses. The understanding of CDs’ luminescence, however, is still far from being comprehensive. The intense photoluminescence can have different origins: molecular mechanisms, oxidation of polyaromatic graphene-like layers, and core-shell interactions of carbonaceous nanoparticles. The citric acid (CA) is one of the most common precursors for CD preparation because of its high biocompatibility, and this review is mainly focused on CA-based CDs. The different parameters that control the synthesis, such as the temperature, the reaction time, and the choice of solvents, were critically described. Particular attention was devoted to the CDs’ optical properties, such as tunable emission and quantum yields, in light of functional applications. The survey of the literature allowed correlating the preparation methods with the structures and the properties of CA-based CDs. Some basic rules to fabricate highly luminescent nanoparticles were selected by the metanalysis of the current literature in the field. In some cases, these findings can be generalized to other types of CDs prepared via liquid phase.

1. Introduction

Since the discovery of fluorescent carbon nanotubes fragments [1], carbon dots (CDs) have been extensively studied as an emerging generation of optoelectronic materials. The reason for such interest is due to both technological and basic science issues. The tunable photoluminescence and the high quantum yield (QY) combined with low-cost preparation [2,3,4,5] are undoubtedly attractive properties for material scientists looking for new light-emitting nanostructures. As a metal-free optical material, CDs are also environmentally friendly and represent a promising alternative to conventional metal-based semiconductors and rare elements. CDs could, therefore, potentially provide outstanding advantages in various fields, such as sensing [6,7], bioimaging [8], catalysis [9], lighting and displaying [10,11], lasering [12,13,14], etc.
At the same time, CDs allow revolutionizing the actual perspective on the classical chemistry classifications. CDs are 0-D emissive spheroidal carbon-based nanostructures with a size smaller than 20 nm [2,5]. The CDs, in fact, stand in between organic (polymers) and inorganic materials (black carbon), macromolecules, and nanoparticle, between bottom-up (polycyclic aromatic compounds) and top-down synthesis (laser ablation of graphene, etc.). However, it is still a challenge to understand the CDs’ correlation between nano-structure and bright fluorescence [15].
The number of publications about CDs has rapidly increased in recent years (see Figure 1a), reflecting the utmost simplicity in preparing the materials but also the galaxy of slightly different results that small changes in the synthesis can produce. An abundance of precursors, in fact, can be used for the preparation of CDs via bottom-up methods, ranging from simple and natural molecules up to complex and expensive compounds [16]. Among them, citric acid (CA), a weak organic acid, is the most popular carbon precursor (see Figure 1b) because of the biocompatibility and the low cost. CA-based CDs show both photoluminescence from blue to red regions [17,18] and extremely high QYs (more than 80%) [6,19].
In this review, we focused on CDs obtained via CA as the main precursor with the purpose to explore the chemistry behind the synthesis. Citric acid, among the various precursors of CDs, is undoubtedly one of the most used, as the abundant scientific literature on the topic demonstrates. Focusing on one precursor allows retrieving a series of data to model the relationships between structure and properties. CA-based CDs can be considered as one of the easiest dots to be obtained; this makes it easier to understand how and why the synthesis affects the CDs properties.
We critically reviewed most of the synthesis parameters which affect the CDs properties, including temperature, reaction time, choice of solvents, etc. Moreover, we focused on the description of CDs’ optical performances and how to improve them. Furthermore, we summarized the theoretical models of CA-based CDs by comparing the different approaches which consider the dots as large molecules or functionalized nanoparticles. Finally, we listed the new exciting applications which are foreseen in the near (bright) future of the CDs.

2. Towards a Retro-Engineering of the CD Structure

The striking possibility of synthesizing such bright fluorescent nanoparticles using simple protocols and cheap precursors is the reason for the broad interest in CDs. However, everything comes at a cost. The multiple carbonization reactions which drive the formation of the emitting nanoparticles are largely uncontrolled, and it appears to be difficult to disentangle such a complex network of processes occurring during the CD formation. Moreover, when hydrothermal, microwave treatment, or thermal degradation are used to synthesize CDs, the chemical reactions which lead to the formation of the carbon nanoparticles are very different from those which typically occur in organic chemistry. A careful retro-engineering process is, therefore, at the basis of real CDs nanotechnology. Bottom-up routes are effective methods to obtain high quality CDs with some organics as precursors. CDs can be synthesized from CA via hydrothermal or solvothermal methods [20,21,22], microwave treatments [23,24], thermal decomposition [25,26], etc. [27,28].

2.1. Carbonization Temperature vs. Optical Properties

It has been experimentally demonstrated that temperature is the determining factor for the carbonization of organic matters when CDs are fabricated. Many researchers have studied the influence of reaction temperature on the optical properties of CDs. We summarized some results in Figure 2, according to the published literature [6,25,29,30,31,32,33,34]. Figure 2a–e show that the QY of the prepared CDs can usually reach a maximum when the hydrothermal temperature (HyT) is set in the range from ~150 to ~200 °C. The correlation between QY and HyT is the main reason why 160, 180, or 200 °C are the typical values selected for the preparation of CDs in many papers. This conclusion is not limited to hydrothermal routes but can be somewhat generalized to other approaches, such as thermal decomposition [25,33] in Figure 2f,g, and microwave treatment [34] in Figure 2h.
Wang et al. [35], Dhenadhayalan et al. [36], and Qu et al. [30] prepared CA-based via thermal decomposition, microwave treatment (aqueous solution), and hydrothermal method, respectively, and compared the optical properties of the products. CDs show similar absorption bands, maximum emission, and QY, suggesting that CDs synthesized from the same raw materials possess similar optical features, even though different preparation ways [37,38,39,40,41,42,43] are carried out if the reaction temperature is kept constant (Figure 3).
It is worth considering the reason for the relationship between temperature and fluorescence intensity. A change of the morphology and the structure at a defined temperature range appears as a reasonable hypothesis. Some researchers [44,45,46,47] have studied the morphologies and the structural characterizations of CA-based CDs at different preparation temperatures. The formation of CDs is divided into different steps: dehydration, polymerization, aromatization, and carbonization. In some cases, transmission electron microscope (TEM) analyses have revealed that CDs change the structure from cross-linked polymer-like to individual nanoparticles when the temperature increases from 150 to 300 °C [48] (see Figure 4a1–d1). Furthermore, X-ray photoelectron spectroscopy (XPS) characterization has shown that the content of graphitic structure C–C/C=C in the CDs increases with the rise of temperature (see Figure 4a2–d2).
In a general model, the CD is made of two key parts, a carbon graphitic core and an amorphous carbon region [50,51,52,53,54] (see Figure 5a–e). A rise of temperature is responsible for the growth of the core and a simultaneous reduction of the amorphous region (Figure 5f) affecting the optical properties. At low temperatures, it is difficult to form a graphitic core, where, therefore, polymeric-like CDs are usually obtained. The photoluminescence of these CDs can be attributed to molecular-state emissive centers [36,48]. On the contrary, the high reaction temperatures tend to fully carbonize the functional groups of the amorphous surface. A larger size of the carbon core generally leads to a shift of photoluminescence towards larger wavelengths because of a quantum confinement effect [55,56], where the optical properties of CDs have shown a size-dependent effect. However, the carbonization of surface groups generally leads to a QY decrease. The highest QY is, therefore, achieved at an intermediate temperature range where core and shell structures coexist. It is necessary to point out that there is no clear temperature boundary to strictly control the formation of molecular-state and aromatic carbon-core domains. The coexistence of both results in multi-type CD emissions, such as size- or excitation-dependent/independent luminescence.

2.2. Reaction Time

The reaction time also affects the optical properties of the related CDs [18,23,25,30,57,58,59]. Figure 6 resumes the plots of QY vs. time measured from different CDs prepared by different groups. Although the syntheses show some differences in terms of reagents and reaction temperatures, the QY reaches a maximum in a time which depends on the reaction temperature and then decreases progressively at longer reaction times. The chemical composition vs. the reaction time shows relevant change, as shown in Figure 4e–g. In short, the rise of reaction time also leads to increased carbonization, which promotes the growth of the carbon core. This trend is similar to the changes occurring as a function of increasing temperatures so that the schematic process shown in Figure 5f is still applicable for increasing reaction times.

2.3. Solvothermal vs. Hydrothermal Method

The selection of solvent is critical to determine the CDs’ properties. For example, blue-emitting CDs were prepared by Zhai et al. [60] through a microwave treatment using CA and ethylenediamine as the precursors and water as the solvent. The same precursors dissolved in formamide, however, allowed preparation of red-emitting CDs [61] with a ~200 nm redshift with respect to the previous protocol. Formamide plays an important role in several admixtures of reagents, enabling fine-tuning of the CDs luminescence. In a typical hydrothermal synthesis, CA and urea form mainly blue-emissive CDs, with a minor emission in the green range [62]. Hola et al. [17], however, obtained CDs with blue, green, and red emissions by simply replacing formamide instead of water during solvothermal carbonization.
Interestingly, the reaction temperature to synthesize CA-based CDs via hydrothermal and solvothermal methods is significantly different. The example is shown in Figure 7. Although the hydrothermal temperature can be typically set in a 150–300 °C range, for instance, the maximum emission of the CDs prepared by this synthesis shows a small shift of around 20 nm [6]. On the contrary, the temperature used for the solvothermal method (dimethylformamide, DMF, as solvent) ranges from 140 to 200 °C. Still, it causes a larger shift in the CDs emission, which can exceed 80 nm [63]. It is, therefore, clear that the use of an organic solvent in the CDs synthesis promotes the carbonization process, leading to major changes in the photophysical properties of the carbon nanoparticles.
Once prepared, the CDs must then be carefully purified to achieve a narrow emission. A lack of purification, in fact, could lead to heterogeneous results which can become very difficult to rationalize. Pan et al. [23], for instance, prepared CDs with CA dissolved in formamide and then dialyzed the products of reaction to separate the CDs with excitation-dependent multicolor emission. On the other hand, another research group [61] which used the same materials and a similar method but a different purification process (solvent-washing) measured an excitation-independent single red photoluminescence.
Kong et al. [64] prepared blue-emitting CDs by hydrothermal treatment with CA and ethylenediamine as a precursor, coupled with dialysis purification. However, Wen et al. [65] were capable of separating the products obtained by the same synthesis through purification via column chromatography, achieving blue- and green- emissive CDs.

2.4. Nitrogen Sources

There are many other parameters in the CDs preparation that have a significant impact on the properties of the products. The presence of a nitrogen source during the CD synthesis has a pivotal role in controlling the emission, however, the reactivity of the different nitrogen sources is not the same and should be carefully taken into consideration. Reckmeier et al. [66] prepared different CDs synthesized from CA using hydrothermal and ammonothermal approaches with aqueous ammonia and supercritical ammonia as the nitrogen source, respectively. They found that the two kinds of CDs are remarkably different in terms of both structure and fluorescence. On the other hand, Sharma et al. [67] obtained CDs by thermal decomposition of CA and urea under N2 and air atmosphere, respectively, however, the two CDs showed similar photophysical properties.
Nitrogen-doping is also a powerful method to increase the QY of CDs photoluminescence. Table 1 compares the QYs of pure CA-based with nitrogen (N)-doped CDs. Interestingly, the N-doping provides higher QY values when specific nitrogen sources are used during the synthesis.
Schneider and co-workers [73] studied the influence of three different nitrogen sources on CA-based CDs: ethylenediamine, hexamethylenetetramine, and triethanolamine. As shown in Figure 8a, ethylenediamine provides the highest QY enhancement, while triethanolamine does not affect the overall fluorescent intensity. Furthermore, when doping is obtained by ethylenediamine, ethanolamine, and tris(hydroxylmethyl)aminomethane as nitrogen sources, the formation of graphitic nitrogen and the enrichment of hydroxyl-groups allow forming CDs with remarkably high QY [44] (Figure 8b).

3. Optical Properties of CA-Based CDs

N-source precursors allow modulating the photophysical properties of CDs in a wide range of emissions. As shown in Table 2, the emission from CA-based CDs can be tuned from the blue to the red with remarkable QYs (>90%). The QY, however, gradually reduces when the characteristic emission is shifted towards higher wavelengths. According to the reports in Table 2, the CA-based CDs can also exhibit near-infrared photoluminescence, especially when synthesized in aprotic polar solvents, such as DMF and formamide [61,75,76].

3.1. Multicolor Photoluminescence

Some specific syntheses allow preparing CA-based CDs with multicolor tunable photoluminescence. This means that minor variations of the synthesis conditions lead to different emissions. By controlling the solvothermal conditions, Miao et al. [63] prepared multiple emissive CDs as shown in Figure 9. The shift of emission from the blue to the red of the visible spectrum is related to the presence of the carboxylic groups and the increasing graphite content in the CDs (Figure 9d). Hola and his co-workers [17] prepared multicolor emissive CDs via solvothermal treatment of CA and urea in a formamide solution. The obtained products were further purified and separated by column chromatography into four fractions showing emissions from blue to red. Herein, however, the redshift of the photoluminescence was attributed to increasing graphitic nitrogen contents, as also confirmed by time-dependent density functional theory (TD-DFT). Interestingly, all the works focused on the multicolor photoluminescence of CDs agree in considering the graphitic structure (derived from the carbonization) as a key factor to control the CDs’ multicolor emission.
While the luminescence appears to be largely controlled by the degree of graphitization of the CD structure, the mechanism controlling the fluorescence is still unclear. This property seems to be affected by many other factors which induce a multicolor fluorescence. Zhu et al. [82], for example, proved that the introduction of metal ions into CA-based CDs structure prepared via magnetic hyperthermia approach leads to a multicolor-fluorescence over a wide range of the visible spectrum. Moreover, some CDs exhibit a long fluorescence lifetime, such as long afterglow performance. CDs fabricated via hydrothermal treatment of CA, acrylamide, and urea exhibit room-temperature phosphorescence with a long lifetime reaching up to 459 ms [83]. Boron is also an important doping agent in the development of CDs featuring ultralong lifetime room temperature phosphorescence. Boron-doped CDs exhibit bright yellow-green afterglow with a remarkable QY of 23.5% and a lifetime of 1.17 s [84].

3.2. Surface Modifications

Surface modification is, in general, an effective strategy to enhance CDs’ optical performances. Usually, the CDs obtained by pyrolysis of CA show a weak fluorescence with a short lifetime. In our previous work [59], it was found that the surface functionalization with 3-aminopropyltriethoxysilane (APTES) reduces the CDs fluorescence quenching due to the formation of a passivation layer. The 3-glycidyloxypropyltrimethoxysilane (GPTMS) is also a possible alternative to modify the surface via an epoxy-amine reaction [22]. In addition to the native luminescence at 430 nm, the GPTMS-grafted CDs show a new emission at 490 nm due to the polyethylene oxide species.
Liu et al. [85] prepared highly emissive solid-state CDs via one-step hydrothermal method with CA as the carbon source and branched poly(ethylenimine) (b-PEI) as the surface passivation agent. The introduction of b-PEI chains can prevent the collisions of CD emissive centers, which further avoids the aggregation-induced quenching.

3.3. External Variables Controlling the Emission

The optical response of phosphors, including CDs, usually are temperature dependent. Kalytchuk et al. [86] prepared nitrogen/sulfur-co-doped CDs by hydrothermal treatment of CA and L-cysteine, which show temperature-dependent absorption, steady-state and transient photoluminescence, and lifetime. The sensitivity to external temperature enables the use of CDs as nano-thermometers featuring good sensitivity [87].
When CDs are used in solution, pH is also a pivotal parameter capable of tuning the emission. CDs made of CA and polyethylenimine (PEI) or 2,3-diaminopyridine (DAP) have shown a reduction of the QYs from ≈ 40–50% to less than 9% when dissolved in an aqueous solution of pH ≈ 1 and to 21% when the pH has been increased to 12 [88]. The peculiar behavior of some CD’s fluorophores formed in situ during CDs carbonization, such as citrazinic acid, has also been recently studied at extreme pH conditions [89,90]. The same fluorophores dissolved in solution have also proved to drastically change their optical properties as a function of the concentration undergoing a transition from monomer to dimer [91]. Concentration-dependent emission is another feature of CDs, which is usually compared with aggregation-induced emission. Wang et al. [37] synthesized a type of concentration-dependent fluorescent CD by microwave treatment of CA and ammonia. The photoluminescence shows a remarkable red-shift when the CDs’ concentration increases from 0.78 to 10.42 mg mL−1 in an aqueous solution.
The abundant surface groups, e.g., –COOH, –NH2, and –OH, contribute to the high solubility of CDs in water. Red-emitting CDs, however, are not prepared by hydrothermal treatment but rather via solvothermal routes employing DMF, ethanol, etc., as solvents [18,63]. Red emitting CDs typically have a low solubility in water with lower QYs [61,75,76]. For example, the 53% QY of red CDs in EtOH drops to 24% when dispersed in water [75].

4. Optical Applications

The excellent optical properties make CDs a promising material to be used in many potential applications, as resumed in Figure 10. The strong absorption endowed by the CDs could be efficiently applied to ultraviolet shielding devices [92]. It is also widely recognized that CDs boost the photocatalytic reactions. CD-based photocatalysts also work well in the near-infrared region where up-conversion is observed [93,94]. CDs-based devices are expected to replace toxic compounds or rare elements in monitors and fluorescent bulbs even if, at the moment, they lack the efficiency requested in the red region. The state of the art for CA-based CDs is multicolor emission with QYs exceeding 50% in the red [65,75,77].
The surface functionalization has already proved to further enhance the CD’s optical properties. Organosilane-linked CDs [95,96], for instance, show both down and up-conversion photoluminescence and exhibit significant multi-photon absorption at room temperature [97]. Although multiphoton excited fluorescence is an emerging research topic, however, it is clear that specifically designed CDs could provide dramatic advances in cutting-edge optical applications such as infrared light detectors, microcavity lasers, etc.
CDs are sensitive to environment changes, such as pH, ions strength, and temperature, which set the conditions for CDs-based optical sensors. For example, CDs with hydroxyl and carboxyl groups on their surface have shown to work well as pH sensors [98]. Currently, CDs have been extensively used to detect iron (III) due to a strong coordination interaction between Fe3+ and the phenolic groups of CDs [6]. Finally, the nontoxicity and the biocompatibility make CDs ideal nanomaterials to develop innovative bioimaging techniques [43,99], biosensors [100], drug-delivery carriers [101], and nano-based antioxidant formulations [102,103].
In the future, new and effective technologies are required for low-cost production of high-quantity CDs.

5. Summary and Outlook

The general trends retrieved from the studies on CA-based CDs can be easily extended to a larger variety of CDs. CDs prepared from folic acid, for instance, show a variable QY which is related to the reaction temperature and has a maximum of around 240 °C [104]. This finding agrees with the correlation between QY and reaction temperature of CA-based CDs.
The choice of the solvent in a solvothermal synthesis is also generally affecting the optical properties and the QY of the products, even for CDs not based on CA. Para-phenylenediamine-based CDs synthesized in water, for instance, show two characteristic emissions, one at 600 nm (red) and a second one at 400 nm (blue). If the reaction solvent is replaced with ethanol, DMF, cyclohexane, or toluene, the red emission increases in intensity, while the blue one decreases [105].
Finally, the increase of graphitic carbon domains generally causes a redshift of photoluminescence in the CDs, although the same effect can be also attributed to other parameters (such as the graphitic nitrogen), especially for CDs not based on CA [106].
Although there is large variability in chemical composition and structure of the different CDs, the large number of works published thus far allows identifying some general trends and common phenomena in the preparation and the optical performances of the CDs.
The carbonization temperature is a key factor for controlling the growth of the carbon core and the density of surface groups in the CDs systems. A carbon core is usually obtained at high temperature, while the surface groups reduce as a function of the temperatures. An intermediate temperature (150–200 °C for CA-based CDs), therefore, should be considered to maximize the QY.
Reaction time plays a role similar to the reaction temperature in the CDs’ synthesis. Extending the carbonization for longer times enhances the degree of carbonization and reduces the functional groups on the surface. In the quest of CDs with high QYs, it is, therefore, of paramount importance to optimize the reaction time to achieve a compromise between the core formation and the chemical composition of the surface. When a solvothermal approach is used, the choice of the solvent affects the CDs fluorescence. Generally speaking, the solvents can also serve as nitrogen sources towards doping of the CDs’ structure.
CA-based CDs show multicolor fluorescence. However, the maximum of the emission is shifted towards longer wavelengths when organic, apolar, and N-containing solvents are used instead of water during the solvothermal synthesis. The nitrogen-doping dramatically contributes to enhancing the QY. This effect, however, depends on the chemical composition of the reagents used for doping. It has been found that, in general, primary amines allows increasing the QY more than secondary and tertiary amines NH2- > –NH– > N≡.
Although the machinery of CD fluorescence is still mostly unexplored, it seems that both amorphous and graphitic carbon atoms could contribute to this effect. The size of the carbon core also influences the emission due to the quantum confinement effect by red shifting the fluorescence redshift.
Concerning the technological advancements, although the CDs show a set of promising features, full optimization of the photophysical properties is required to promote the development of innovative functional devices.

Author Contributions

The original draft was written by J.R. and L.M., supervision was performed by L.M. and P.I., the final version was edited by all authors. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Italian Ministry of University and Research (MUR) through project PRIN2017 “CANDL2”, grant number 2017W75RAE.

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. 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. [Google Scholar] [CrossRef]
  2. Zhu, S.; Song, Y.; Zhao, X.; Shao, J.; Zhang, J.; Yang, B. The photoluminescence mechanism in carbon dots (graphene quantum dots, carbon nanodots, and polymer dots): Current state and future perspective. Nano Res. 2015, 8, 355. [Google Scholar] [CrossRef]
  3. Cao, L.; Wang, X.; Meziani, M.J.; Lu, F.; Wang, H.; Luo, P.G.; Lin, Y.; Harruff, B.A.; Veca, L.M.; Murray, D. Carbon dots for multiphoton bioimaging. J. Am. Chem. Soc. 2007, 129, 11318. [Google Scholar] [CrossRef] [Green Version]
  4. 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. [Google Scholar] [CrossRef] [PubMed]
  5. Malfatti, L.; Innocenzi, P. Sol-gel chemistry for carbon dots. Chem. Rec. 2018, 18, 1192. [Google Scholar] [CrossRef] [PubMed]
  6. Zhu, S.; Meng, Q.; Wang, L.; Zhang, J.; Song, Y.; Jin, H.; Zhang, K.; Sun, H.; Wang, H.; Yang, B. Highly photoluminescent carbon dots for multicolor patterning, sensors, and bioimaging. Angew. Chem. Int. Ed. 2013, 52, 3953. [Google Scholar] [CrossRef] [PubMed]
  7. Yang, D.; Ren, J.; Li, J.; Wang, Q.; Wang, Q.; Xie, Z.; Qu, X. Construction of bi-layer biluminophore fast-responding pressure sensitive coating for non-contact unsteady aerodynamic testing. Polym. Test. 2019, 77, 105922. [Google Scholar] [CrossRef]
  8. Luo, P.G.; Sahu, S.; Yang, S.; Sonkar, S.K.; Wang, J.; Wang, H.; LeCroy, G.E.; Cao, L.; Sun, Y. Carbon “quantum” dots for optical bioimaging. J. Mater. Chem. 2013, 1, 2116. [Google Scholar] [CrossRef]
  9. Hutton, G.A.; Martindale, B.C.; Reisner, E. Carbon dots as photosensitisers for solar-driven catalysis. Chem. Soc. Rev. 2017, 46, 6111. [Google Scholar] [CrossRef] [Green Version]
  10. Wang, F.; Chen, Y.; Liu, C.; Ma, D. White light-emitting devices based on carbon dots’ electroluminescence. Chem. Commun. 2011, 47, 3502. [Google Scholar] [CrossRef]
  11. Ren, J.; Sun, J.; Sun, X.; Song, R.; Xie, Z.; Zhou, S. Precisely controlled up/down-conversion liquid and solid state photoluminescence of carbon dots. Adv. Opt. Mater. 2018, 6, 1800115. [Google Scholar] [CrossRef]
  12. Zhang, W.; Zhu, H.; Yu, S.; Yang, H. Observation of lasing emission from carbon nanodots in organic solvents. Adv. Mater. 2012, 24, 2263. [Google Scholar] [CrossRef] [PubMed]
  13. Ren, J.; Sun, X.; Wang, Y.; Song, R.; Xie, Z.; Zhou, S.; Chen, P. Controllable photoluminescent and nonlinear optical properties of polymerizable carbon dots and their arbitrary copolymerized gel glasses. Adv. Opt. Mater. 2018, 6, 1701273. [Google Scholar] [CrossRef]
  14. Han, Z.; Ni, Y.; Ren, J.; Zhang, W.; Wang, Y.; Xie, Z.; Zhou, S.; Yu, S.F. Highly efficient and ultra-narrow bandwidth orange emissive carbon dots for microcavity lasers. Nanoscale 2019, 11, 11577. [Google Scholar] [CrossRef]
  15. Ren, J.; Stagi, L.; Innocenzi, P. Fluorescent carbon dots in solid-state: From nanostructures to functional devices. Prog. Solid State Chem. 2020, 100295. [Google Scholar] [CrossRef]
  16. Semeniuk, M.; Yi, Z.; Poursorkhabi, V.; Tjong, J.; Jaffer, S.; Lu, Z.-H.; Sain, M. Future perspectives and review on organic carbon dots in electronic applications. ACS Nano 2019, 13, 6224. [Google Scholar] [CrossRef]
  17. Hola, K.; Sudolska, M.; Kalytchuk, S.; Nachtigallova, D.; Rogach, A.L.; Otyepka, M.; Zboril, R. Graphitic nitrogen triggers red fluorescence in carbon dots. ACS Nano 2017, 11, 12402. [Google Scholar] [CrossRef]
  18. Yuan, F.; Wang, Z.; Li, X.; Li, Y.; Tan, Z.; Fan, L.; Yang, S. Bright multicolor bandgap fluorescent carbon quantum dots for electroluminescent light-emitting diodes. Adv. Mater. 2017, 29, 1604436. [Google Scholar] [CrossRef]
  19. Xie, Z.; Wang, F.; Liu, C. Organic-inorganic hybrid functional carbon dot gel glasses. Adv. Mater. 2012, 24, 1716. [Google Scholar] [CrossRef]
  20. Chahal, S.; Yousefi, N.; Tufenkji, N. Green synthesis of high quantum yield carbon dots from phenylalanine and citric acid: Role of stoichiometry and nitrogen doping. ACS Sustain. Chem. Eng. 2020, 8, 5566–5575. [Google Scholar] [CrossRef]
  21. Qu, S.; Zhou, D.; Li, D.; Ji, W.; Jing, P.; Han, D.; Liu, L.; Zeng, H.; Shen, D. Toward efficient orange emissive carbon nanodots through conjugated sp2-domain controlling and surface charges engineering. Adv. Mater. 2016, 28, 3516. [Google Scholar] [CrossRef] [PubMed]
  22. Suzuki, K.; Malfatti, L.; Takahashi, M.; Carboni, D.; Messina, F.; Tokudome, Y.; Takemoto, M.; Innocenzi, P. Design of carbon dots photoluminescence through organo-functional silane grafting for solid-state emitting devices. Sci. Rep. 2017, 7, 5469. [Google Scholar] [CrossRef] [PubMed]
  23. Pan, L.; Sun, S.; Zhang, A.; Jiang, K.; Zhang, L.; Dong, C.; Huang, Q.; Wu, A.; Lin, H. Truly fluorescent excitation-dependent carbon dots and their applications in multicolor cellular imaging and multidimensional sensing. Adv. Mater. 2015, 27, 7782. [Google Scholar] [CrossRef] [PubMed]
  24. Carbonaro, C.M.; Chiriu, D.; Stagi, L.; Casula, M.F.; Thakkar, S.V.; Malfatti, L.; Suzuki, K.; Ricci, P.C.; Corpino, R. Carbon dots in water and mesoporous matrix: Chasing the origin of their photoluminescence. J. Phys. Chem. 2018, 122, 25638. [Google Scholar] [CrossRef]
  25. Yin, C.; Fan, Y.; Yang, X.; Zhou, X. Highly efficient synthesis of N-doped carbon dots with excellent stability through pyrolysis method. J. Mater. Sci. 2019, 54, 9372. [Google Scholar] [CrossRef]
  26. Kasprzyk, W.; Bednarz, S.; Żmudzki, P.; Galica, M.; Bogdał, D. Novel efficient fluorophores synthesized from citric acid. RSC Adv. 2015, 5, 34795. [Google Scholar] [CrossRef]
  27. Mura, S.; Ludmerczki, R.; Stagi, L.; Garroni, S.; Carbonaro, C.M.; Ricci, P.C.; Casula, M.F.; Malfatti, L.; Innocenzi, P. Integrating sol-gel and carbon dots chemistry for the fabrication of fluorescent hybrid organic-inorganic films. Sci. Rep. 2020, 10, 4770. [Google Scholar] [CrossRef] [Green Version]
  28. Das, A.; Gude, V.; Roy, D.; Chatterjee, T.; De, C.K.; Mandal, P.K. On the molecular origin of photoluminescence of nonblinking carbon dot. J. Phys. Chem. 2017, 121, 9634. [Google Scholar] [CrossRef]
  29. Liao, S.; Zhao, X.; Zhu, F.; Chen, M.; Wu, Z.; Yang, H.; Chen, X. Novel S, N-doped carbon quantum dot-based “off-on” fluorescent sensor for silver ion and cysteine. Talanta 2018, 180, 300. [Google Scholar] [CrossRef]
  30. Qu, D.; Zheng, M.; Zhang, L.; Zhao, H.; Xie, Z.; Jing, X.; Haddad, R.E.; Fan, H.; Sun, Z. Formation mechanism and optimization of highly luminescent N-doped graphene quantum dots. Sci. Rep. 2014, 4, 5294. [Google Scholar] [CrossRef]
  31. Li, X.; Zhang, S.; Kulinich, S.A.; Liu, Y.; Zeng, H. Engineering surface states of carbon dots to achieve controllable luminescence for solid-luminescent composites and sensitive Be2+ detection. Sci. Rep. 2014, 4, 4976. [Google Scholar] [CrossRef]
  32. Kundu, A.; Lee, J.; Park, B.; Ray, C.; Sankar, K.V.; Kim, W.S.; Lee, S.H.; Cho, I.-J.; Jun, S.C. Facile approach to synthesize highly fluorescent multicolor emissive carbon dots via surface functionalization for cellular imaging. J. Colloid Interface Sci. 2018, 513, 505. [Google Scholar] [CrossRef] [PubMed]
  33. Wang, S.; Chen, Z.; Cole, I.; Li, Q. Structural evolution of graphene quantum dots during thermal decomposition of citric acid and the corresponding photoluminescence. Carbon 2015, 82, 304. [Google Scholar] [CrossRef] [Green Version]
  34. Xiao, Q.; Liang, Y.; Zhu, F.; Lu, S.; Huang, S. Microwave-assisted one-pot synthesis of highly luminescent N-doped carbon dots for cellular imaging and multi-ion probing. Microchim. Acta 2017, 184, 2429. [Google Scholar] [CrossRef]
  35. Wang, J.; Wei, J.; Su, S.; Qiu, J. Novel fluorescence resonance energy transfer optical sensors for vitamin B12 detection using thermally reduced carbon dots. New J. Chem. 2015, 39, 501. [Google Scholar] [CrossRef]
  36. Dhenadhayalan, N.; Lin, K.C.; Suresh, R.; Ramamurthy, P. Unravelling the multiple emissive states in citric-acid-derived carbon dots. J. Phys. Chem. 2016, 120, 1252. [Google Scholar] [CrossRef]
  37. Wang, C.; Hu, T.; Wen, Z.; Zhou, J.; Wang, X.; Wu, Q.; Wang, C. Concentration-dependent color tunability of nitrogen-doped carbon dots and their application for iron (III) detection and multicolor bioimaging. J. Colloid Interface Sci. 2018, 521, 33. [Google Scholar] [CrossRef]
  38. Zhang, Y.; Cui, P.; Zhang, F.; Feng, X.; Wang, Y.; Yang, Y.; Liu, X. Fluorescent probes for “off-on” highly sensitive detection of Hg2+ and L-cysteine based on nitrogen-doped carbon dots. Talanta 2016, 152, 288. [Google Scholar] [CrossRef]
  39. Du, F.; Zeng, F.; Ming, Y.; Wu, S. Carbon dots-based fluorescent probes for sensitive and selective detection of iodide. Microchim. Acta 2013, 180, 453. [Google Scholar] [CrossRef]
  40. Barati, A.; Shamsipur, M.; Abdollahi, H. Hemoglobin detection using carbon dots as a fluorescence probe. Biosens. Bioelectron. 2015, 71, 470. [Google Scholar] [CrossRef]
  41. Feng, T.; Ai, X.; An, G.; Yang, P.; Zhao, Y. Charge-convertible carbon dots for imaging-guided drug delivery with enhanced in vivo cancer therapeutic efficiency. ACS Nano 2016, 10, 4410. [Google Scholar] [CrossRef] [PubMed]
  42. Qu, D.; Zheng, M.; Li, J.; Xie, Z.; Sun, Z. Tailoring color emissions from N-doped graphene quantum dots for bioimaging applications. Light Sci. Appl. 2015, 4, 364. [Google Scholar] [CrossRef]
  43. Wang, J.; Zhang, P.; Huang, C.; Liu, G.; Leung, K.C.F.; Wáng, Y.X.N.J. High performance photoluminescent carbon dots for in vitro and in vivo bioimaging: Effect of nitrogen doping ratios. Langmuir 2015, 31, 8063. [Google Scholar] [CrossRef] [PubMed]
  44. Xiong, Y.; Schneider, J.; Reckmeier, C.J.; Huang, H.; Kasák, P.; Rogach, A.L. Carbonization conditions influence the emission characteristics and the stability against photobleaching of nitrogen doped carbon dots. Nanoscale 2017, 9, 11730. [Google Scholar] [CrossRef] [PubMed]
  45. Ogi, T.; Aishima, K.; Permatasari, F.A.; Iskandar, F.; Tanabe, E.; Okuyama, K. Kinetics of nitrogen-doped carbon dot formation via hydrothermal synthesis. New J. Chem. 2016, 40, 5555. [Google Scholar] [CrossRef]
  46. 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. [Google Scholar] [CrossRef]
  47. Zhou, Y.; Desserre, A.; Sharma, S.K.; Li, S.; Marksberry, M.H.; Chusuei, C.C.; Blackwelder, P.L.; Leblanc, R.M. Gel-like carbon dots: Characterization and their potential applications. ChemPhysChem 2017, 18, 890. [Google Scholar] [CrossRef]
  48. Shamsipur, M.; Barati, A.; Taherpour, A.A.; Jamshidi, M. Resolving the multiple emission centers in carbon dots: From fluorophore molecular states to aromatic domain states and carbon-core states. J. Phys. Chem. Lett. 2018, 9, 4189. [Google Scholar] [CrossRef]
  49. Shang, W.; Cai, T.; Zhang, Y.; Liu, D.; Liu, S. Facile one pot pyrolysis synthesis of carbon quantum dots and graphene oxide nanomaterials: All carbon hybrids as eco-environmental lubricants for low friction and remarkable wear-resistance. Tribol. Int. 2018, 118, 373. [Google Scholar] [CrossRef]
  50. Dong, Y.; Wang, R.; Li, G.; Chen, C.; Chi, Y.; Chen, G. Polyamine-functionalized carbon quantum dots as fluorescent probes for selective and sensitive detection of copper ions. Anal. Chem. 2012, 84, 6220. [Google Scholar] [CrossRef]
  51. Deng, L.; Wang, X.; Kuang, Y.; Wang, C.; Luo, L.; Wang, F.; Sun, X. Development of hydrophilicity gradient ultracentrifugation method for photoluminescence investigation of separated non-sedimental carbon dots. Nano Res. 2015, 8, 2810. [Google Scholar] [CrossRef]
  52. Bhattacharyya, S.; Ehrat, F.; Urban, P.; Teves, R.; Wyrwich, R.; Döblinger, M.; Feldmann, J.; Urban, A.S.; Stolarczyk, J.K. Effect of nitrogen atom positioning on the trade-off between emissive and photocatalytic properties of carbon dots. Nat. Commun. 2017, 8, 1401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Kwon, W.; Lee, G.; Do, S.; Joo, T.; Rhee, S.W. Size-controlled soft-template synthesis of carbon nanodots toward versatile photoactive materials. Small 2014, 10, 506. [Google Scholar] [CrossRef] [PubMed]
  54. Hou, X.; Hu, Y.; Wang, P.; Yang, L.; Al Awak, M.M.; Tang, Y.; Twara, F.K.; Qian, H.; Sun, Y. Modified facile synthesis for quantitatively fluorescent carbon dots. Carbon 2017, 122, 389. [Google Scholar] [CrossRef] [PubMed]
  55. Zhu, J.; Bai, X.; Chen, X.; Shao, H.; Zhai, Y.; Pan, G.; Zhang, H.; Ushakova, E.V.; Zhang, Y.; Song, H. Spectrally tunable solid state fluorescence and room-temperature phosphorescence of carbon dots synthesized via seeded growth method. Adv. Opt. Mater. 2019, 7, 1801599. [Google Scholar] [CrossRef]
  56. Wang, B.; Yu, J.; Sui, L.; Zhu, S.; Tang, Z.; Yang, B.; Lu, S. Rational design of multi-color-emissive carbon dots in a single reaction system by hydrothermal. Adv. Sci. 2020, 2001453. [Google Scholar] [CrossRef]
  57. Gao, F.; Ma, S.; Li, J.; Dai, K.; Xiao, X.; Zhao, D.; Gong, W. Rational design of high quality citric acid-derived carbon dots by selecting efficient chemical structure motifs. Carbon 2017, 112, 131. [Google Scholar] [CrossRef]
  58. Wang, F.; Pang, S.; Wang, L.; Li, Q.; Kreiter, M.; Liu, C.Y. One-step synthesis of highly luminescent carbon dots in noncoordinating solvents. Chem. Mater. 2010, 22, 4528. [Google Scholar] [CrossRef]
  59. Ludmerczki, R.; Mura, S.; Carbonaro, C.M.; Mandity, I.M.; Carraro, M.; Senes, N.; Garroni, S.; Granozzi, G.; Calvillo, L.; Marras, S. Carbon dots from citric acid and its intermediates formed by thermal decomposition. Chem. Eur. J. 2019, 25, 11963. [Google Scholar] [CrossRef]
  60. Zhai, Y.; Zhu, Z.; Zhu, C.; Ren, J.; Wang, E.; Dong, S. Multifunctional water-soluble luminescent carbon dots for imaging and Hg2+ sensing. J. Mater. Chem. 2014, 2, 6995. [Google Scholar] [CrossRef]
  61. Sun, S.; Zhang, L.; Jiang, K.; Wu, A.; Lin, H. Toward high-efficient red emissive carbon dots: Facile preparation, unique properties, and applications as multifunctional theranostic agents. Chem. Mater. 2016, 28, 8659. [Google Scholar] [CrossRef]
  62. Zhang, Y.; He, Y.; Cui, P.; Feng, X.; Chen, L.; Yang, Y.; Liu, X. Water-soluble, nitrogen-doped fluorescent carbon dots for highly sensitive and selective detection of Hg2+ in aqueous solution. RSC Adv. 2015, 5, 40393. [Google Scholar] [CrossRef]
  63. Miao, X.; Qu, D.; Yang, D.; Nie, B.; Zhao, Y.; Fan, H.; Sun, Z. Synthesis of carbon dots with multiple color emission by controlled graphitization and surface functionalization. Adv. Mater. 2018, 30, 1704740. [Google Scholar] [CrossRef] [PubMed]
  64. Kong, T.; Hao, L.; Wei, Y.; Cai, X.; Zhu, B. Doxorubicin conjugated carbon dots as a drug delivery system for human breast cancer therapy. Cell Prolif. 2018, 51, e12488. [Google Scholar] [CrossRef] [Green Version]
  65. Wen, Z.; Yin, X. Excitation-independent carbon dots, from photoluminescence mechanism to single-color application. RSC Adv. 2016, 6, 27829. [Google Scholar] [CrossRef]
  66. Reckmeier, C.J.; Schneider, J.; Xiong, Y.; Häusler, J.; Kasák, P.; Schnick, W.; Rogach, A.L. Aggregated molecular fluorophores in the ammonothermal synthesis of carbon dots. Chem. Mater. 2017, 29, 10352. [Google Scholar] [CrossRef] [Green Version]
  67. Sharma, A.; Gadly, T.; Gupta, A.; Ballal, A.; Ghosh, S.K.; Kumbhakar, M. Origin of excitation dependent fluorescence in carbon nanodots. J. Phys. Chem. Lett. 2016, 7, 3695. [Google Scholar] [CrossRef]
  68. Sun, C.; Zhang, Y.; Sun, K.; Reckmeier, C.; Zhang, T.; Zhang, X.; Zhao, J.; Wu, C.; William, W.Y.; Rogach, A.L. Combination of carbon dot and polymer dot phosphors for white light-emitting diodes. Nanoscale 2015, 7, 12045. [Google Scholar] [CrossRef]
  69. 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. [Google Scholar] [CrossRef]
  70. Zuo, P.; Liu, J.; Guo, H.; Wang, C.; Liu, H.; Zhang, Z.; Liu, Q. Multifunctional N, S co-doped carbon dots for sensitive probing of temperature, ferric ion, and methotrexate. Anal. Bioanal. Chem. 2019, 411, 1647. [Google Scholar] [CrossRef]
  71. Liu, Y.; Zhou, L.; Li, Y.; Deng, R.; Zhang, H. Highly fluorescent nitrogen-doped carbon dots with excellent thermal and photo stability applied as invisible ink for loading important information and anti-counterfeiting. Nanoscale 2017, 9, 491. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Wang, H.; Lu, Q.; Hou, Y.; Liu, Y.; Zhang, Y. High fluorescence S, N co-doped carbon dots as an ultra-sensitive fluorescent probe for the determination of uric acid. Talanta 2016, 155, 62. [Google Scholar] [CrossRef] [PubMed]
  73. Schneider, J.; Reckmeier, C.J.; Xiong, Y.; Von Seckendorff, M.; Susha, A.S.; Kasák, P.; Rogach, A.L. Molecular fluorescence in citric acid-based carbon dots. J. Phys. Chem. 2017, 121, 2014. [Google Scholar] [CrossRef]
  74. Zhang, Y.; Liu, X.; Fan, Y.; Guo, X.; Zhou, L.; Lv, Y.; Lin, J. One-step microwave synthesis of N-doped hydroxyl-functionalized carbon dots with ultra-high fluorescence quantum yields. Nanoscale 2016, 8, 15281. [Google Scholar] [CrossRef]
  75. Ding, H.; Wei, J.; Zhong, N.; Gao, Q.; Xiong, H. Highly efficient red-emitting carbon dots with gram-scale yield for bioimaging. Langmuir 2017, 33, 12635. [Google Scholar] [CrossRef]
  76. Li, D.; Jing, P.; Sun, L.; An, Y.; Shan, X.; Lu, X.; Zhou, D.; Han, D.; Shen, D.; Zhai, Y. Near-infrared excitation/emission and multiphoton-induced fluorescence of carbon dots. Adv. Mater. 2018, 30, 1705913. [Google Scholar] [CrossRef]
  77. Anjana, R.; Devi, J.A.; Jayasree, M.; Aparna, R.; Aswathy, B.; Praveen, G.; Lekha, G.; Sony, G. S, N-doped carbon dots as a fluorescent probe for bilirubin. Microchim. Acta 2018, 185, 11. [Google Scholar] [CrossRef]
  78. Zhang, W.; Shi, L.; Liu, Y.; Meng, X.; Xu, H.; Xu, Y.; Liu, B.; Fang, X.; Li, H.; Ding, T. Supramolecular interactions via hydrogen bonding contributing to citric-acid derived carbon dots with high quantum yield and sensitive photoluminescence. RSC Adv. 2017, 7, 20345. [Google Scholar] [CrossRef] [Green Version]
  79. Liu, Q.; Zhang, N.; Shi, H.; Ji, W.; Guo, X.; Yuan, W.; Hu, Q. One-step microwave synthesis of carbon dots for highly sensitive and selective detection of copper ions in aqueous solution. New J. Chem. 2018, 42, 3097. [Google Scholar] [CrossRef]
  80. Hou, J.; Wang, W.; Zhou, T.; Wang, B.; Li, H.; Ding, L. Synthesis and formation mechanistic investigation of nitrogen-doped carbon dots with high quantum yields and yellowish-green fluorescence. Nanoscale 2016, 8, 11185. [Google Scholar] [CrossRef]
  81. Cheng, J.; Wang, C.; Zhang, Y.; Yang, S.; Chen, S. Zinc ion-doped carbon dots with strong yellow photoluminescence. RSC Adv. 2016, 6, 37189. [Google Scholar] [CrossRef]
  82. Zhu, Z.; Cheng, R.; Ling, L.; Li, Q.; Chen, S. Rapid and large-scale production of multi-fluorescence carbon dots via magnetic hyperthermia method. Angew. Chem. Int. Ed. 2019, 59, 3099. [Google Scholar] [CrossRef] [PubMed]
  83. Li, H.; Ye, S.; Guo, J.; Kong, J.; Song, J.; Kang, Z. The design of room-temperature-phosphorescent carbon dots and their application as a security ink. J. Mater. Chem. 2019, 7, 10605. [Google Scholar] [CrossRef]
  84. Xie, Z.; Zheng, M. Colour-tunable ultralong-lifetime room temperature phosphorescence with external heavy-atom effect in boron-doped carbon dots. Chem. Eng. J. 2020, 127647. [Google Scholar] [CrossRef]
  85. Liu, E.; Li, D.; Zhou, X.; Zhou, G.; Xiao, H.; Zhou, D.; Tian, P.; Guo, R.; Qu, S. Highly emissive carbon dots in solid state and their applications in light-emitting devices and visible light communication. ACS Sustain. Chem. Eng. 2019, 7, 9301. [Google Scholar] [CrossRef]
  86. Kalytchuk, S.; Poláková, K.I.; Wang, Y.; Froning, J.P.; Cepe, K.; Rogach, A.L.; Zbořil, R. Carbon dot nanothermometry: Intracellular photoluminescence lifetime thermal sensing. ACS Nano 2017, 11, 1432. [Google Scholar] [CrossRef]
  87. Mohammed, L.J.; Omer, K.M. Carbon dots as new generation materials for nanothermometer. Nanoscale Res. Lett. 2020, 15, 182. [Google Scholar] [CrossRef]
  88. Meierhofer, F.; Dissinger, F.; Weigert, F.; Jungclaus, J.; Müller-Caspary, K.; Waldvogel, S.R.; Resch-Genger, U.; Voss, T. Citric acid based carbon dots with amine type stabilizers: pH-specific luminescence and quantum yield characteristics. J. Phys. Chem. 2020, 124, 8894. [Google Scholar] [CrossRef]
  89. Stagi, L.; Mura, S.; Malfatti, L.; Carbonaro, C.M.; Ricci, P.C.; Porcu, S.; Secci, F.; Innocenzi, P. Anomalous optical properties of citrazinic acid under extreme pH conditions. ACS Omega 2020. [Google Scholar] [CrossRef]
  90. Mura, S.; Stagi, L.; Ludmerczki, R.; Malfatti, L.; Innocenzi, P. Reversible aggregation of molecular-like fluorophores driven by extreme pH in carbon dots. Materials 2020, 13, 3654. [Google Scholar] [CrossRef]
  91. Mura, S.; Stagi, L.; Malfatti, L.; Carbonaro, C.M.; Ludmerczki, R.; Innocenzi, P. Modulating the optical properties of citrazinic acid through the monomer-to-dimer transformation. J. Phys. Chem. 2019, 124, 197. [Google Scholar] [CrossRef] [PubMed]
  92. Xie, Z.; Du, Q.; Wu, Y.; Hao, X.; Liu, C. Full-band UV shielding and highly daylight luminescent silane-functionalized graphene quantum dot nanofluids and their arbitrary polymerized hybrid gel glasses. J. Mater. Chem. 2016, 4, 9879. [Google Scholar] [CrossRef]
  93. Wang, H.; Sun, P.; Cong, S.; Wu, J.; Gao, L.; Wang, Y.; Dai, X.; Yi, Q.; Zou, G. Nitrogen-doped carbon dots for “green” quantum dot solar cells. Nanoscale Res. Lett. 2016, 11, 27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Zhang, Y.; Park, M.; Kim, H.Y.; Ding, B.; Park, S.-J. A facile ultrasonic-assisted fabrication of nitrogen-doped carbon dots/BiOBr up-conversion nanocomposites for visible light photocatalytic enhancements. Sci. Rep. 2017, 7, 45086. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Liu, J.; Ren, J.; Xie, Z.; Guan, B.; Wang, J.; Ikeda, T.; Jiang, L. Multi-functional organosilane-polymerized carbon dot inverse opals. Nanoscale 2018, 10, 4642. [Google Scholar] [CrossRef]
  96. Liu, J.; Xie, Z.; Shang, Y.; Ren, J.; Hu, R.; Guan, B.; Wang, J.; Ikeda, T.; Jiang, L. Lyophilic but nonwettable organosilane-polymerized carbon dots inverse opals with closed-cell structure. ACS Appl. Mater. Interfaces 2018, 10, 6701. [Google Scholar] [CrossRef]
  97. Zhang, W.; Ni, Y.; Xu, X.; Lu, W.; Ren, P.; Yan, P.; Siu, C.K.; Ruan, S.; Yu, S.F. Realization of multiphoton lasing from carbon nanodot microcavities. Nanoscale 2017, 9, 5957. [Google Scholar] [CrossRef]
  98. Shangguan, J.; He, D.; He, X.; Wang, K.; Xu, F.; Liu, J.; Tang, J.; Yang, X.; Huang, J. Label-free carbon-dots-based ratiometric fluorescence pH nanoprobes for intracellular pH sensing. Anal. Chem. 2016, 88, 7837. [Google Scholar] [CrossRef]
  99. Chen, Y.; Sun, X.; Wang, X.; Pan, W.; Yu, G.; Wang, J. Carbon dots with red emission for bioimaging of fungal cells and detecting Hg2+ and ziram in aqueous solution. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 118230. [Google Scholar] [CrossRef]
  100. Wang, X.; Yan, P.; Kerns, P.; Suib, S.L.; Loew, L.; Zhao, J. Voltage-dependent photoluminescence of carbon dots. J. Electrochem. Soc. 2020, 167, 147515. [Google Scholar] [CrossRef]
  101. Zeng, Q.; Shao, D.; He, X.; Ren, Z.; Ji, W.; Shan, C.; Qu, S.; Li, J.; Chen, L.; Li, Q. Carbon dots as a trackable drug delivery carrier for localized cancer therapy in vivo. J. Mater. Chem. 2016, 4, 5119. [Google Scholar] [CrossRef] [PubMed]
  102. Ji, Z.; Yin, Z.; Jia, Z.; Wei, J. Carbon nanodots derived from urea and citric acid in living cells: Cellular uptake and antioxidation effect. Langmuir 2020, 36, 8632. [Google Scholar] [CrossRef] [PubMed]
  103. Wang, H.; Zhang, M.; Ma, Y.; Wang, B.; Huang, H.; Liu, Y.; Shao, M.; Kang, Z. Carbon dots derived from citric acid and glutathione as a highly efficient intracellular reactive oxygen species scavenger for alleviating the lipopolysaccharide-induced inflammation in macrophages. ACS Appl. Mater. Interfaces 2020, 12, 41088. [Google Scholar] [CrossRef] [PubMed]
  104. Liu, H.; Li, Z.; Sun, Y.; Geng, X.; Hu, Y.; Meng, H.; Ge, J.; Qu, L. Synthesis of luminescent carbon dots with ultrahigh quantum yield and inherent folate receptor-positive cancer cell targetability. Sci. Rep. 2018, 8, 1086. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  105. Zhang, T.; Zhu, J.; Zhai, Y.; Wang, H.; Bai, X.; Dong, B.; Wang, H.; Song, H. A novel mechanism for red emission carbon dots: Hydrogen bond dominated molecular states emission. Nanoscale 2017, 9, 13042. [Google Scholar] [CrossRef] [PubMed]
  106. Ding, H.; Wei, J.S.; Zhang, P.; Zhou, Z.Y.; Gao, Q.Y.; Xiong, H.M. Solvent-controlled synthesis of highly luminescent carbon dots with a wide color gamut and narrowed emission peak widths. Small 2018, 14, 1800612. [Google Scholar] [CrossRef]
Figure 1. (a) Publications since 2004 retrieved by searching “carbon dots” or “carbon nanodots” or “graphene quantum dots” on Web of Science in November 2020; (b) publications since 2010 by refining the obtained result with “citric acid”.
Figure 1. (a) Publications since 2004 retrieved by searching “carbon dots” or “carbon nanodots” or “graphene quantum dots” on Web of Science in November 2020; (b) publications since 2010 by refining the obtained result with “citric acid”.
Carbon 07 00002 g001
Figure 2. Temperature vs. quantum yields of carbon dots (CDs) synthesized from citric acid (CA) by (ae) hydrothermal method, (f,g) thermal decomposition, and (h) microwave treatment. The data in (ah) are organized according to the results in Ref. [29], Ref. [30], Ref. [31], Ref. [32], Ref. [6], Ref. [33] Ref. [25], and Ref. [34], respectively.
Figure 2. Temperature vs. quantum yields of carbon dots (CDs) synthesized from citric acid (CA) by (ae) hydrothermal method, (f,g) thermal decomposition, and (h) microwave treatment. The data in (ah) are organized according to the results in Ref. [29], Ref. [30], Ref. [31], Ref. [32], Ref. [6], Ref. [33] Ref. [25], and Ref. [34], respectively.
Carbon 07 00002 g002
Figure 3. The optical properties of different CDs. CDs in (a1, Ref. [37]) and (a2, Ref. [38]) are synthesized from CA and ammonia by microwave method and hydrothermal treatment, respectively. CDs in (b1, Ref. [39]) and (b2, Ref. [40]) are synthesized from CA and ethylenediamine by microwave method and hydrothermal treatment, respectively. CDs in (c1, Ref. [41]), (c2, Ref. [42]), and (c3, Ref. [43]) are synthesized from CA and diethylenetriamine by thermal decomposition and hydrothermal treatment, respectively. Reproduced from Ref. [37,38,39,40,41,42,43] with permission of Copyright (2016, 2018) American Chemical Society, Copyright (2015, 2016, 2018) Elsevier, Copyright (2013) Springer, and Copyright (2015) Nature Publishing Group, respectively.
Figure 3. The optical properties of different CDs. CDs in (a1, Ref. [37]) and (a2, Ref. [38]) are synthesized from CA and ammonia by microwave method and hydrothermal treatment, respectively. CDs in (b1, Ref. [39]) and (b2, Ref. [40]) are synthesized from CA and ethylenediamine by microwave method and hydrothermal treatment, respectively. CDs in (c1, Ref. [41]), (c2, Ref. [42]), and (c3, Ref. [43]) are synthesized from CA and diethylenetriamine by thermal decomposition and hydrothermal treatment, respectively. Reproduced from Ref. [37,38,39,40,41,42,43] with permission of Copyright (2016, 2018) American Chemical Society, Copyright (2015, 2016, 2018) Elsevier, Copyright (2013) Springer, and Copyright (2015) Nature Publishing Group, respectively.
Carbon 07 00002 g003
Figure 4. (a1d2, Ref. [48]) TEM images and high-resolution XPS spectra of C1s of CDs prepared at different temperatures. (eg, Ref. [49]) The high-resolution XPS spectra of C1s of CDs were prepared with different reaction times. Adopted from Ref. [48,49] with permission of Copyright (2018) American Chemical Society and Copyright (2018) Elsevier.
Figure 4. (a1d2, Ref. [48]) TEM images and high-resolution XPS spectra of C1s of CDs prepared at different temperatures. (eg, Ref. [49]) The high-resolution XPS spectra of C1s of CDs were prepared with different reaction times. Adopted from Ref. [48,49] with permission of Copyright (2018) American Chemical Society and Copyright (2018) Elsevier.
Carbon 07 00002 g004
Figure 5. (ae, Ref. [50,51,52,53,54]) Different models of CD structure diagrams. (f) Structure changes of CDs with the increase of temperature during the carbonization step. Reproduced from Ref. [50,51,52,53,54] with permission of Copyright (2012) American Chemical Society, Copyright (2015) Springer, Copyright (2017) Nature Publishing Group, Copyright (2014) John Wiley and Son, and Copyright (2017) Elsevier, respectively.
Figure 5. (ae, Ref. [50,51,52,53,54]) Different models of CD structure diagrams. (f) Structure changes of CDs with the increase of temperature during the carbonization step. Reproduced from Ref. [50,51,52,53,54] with permission of Copyright (2012) American Chemical Society, Copyright (2015) Springer, Copyright (2017) Nature Publishing Group, Copyright (2014) John Wiley and Son, and Copyright (2017) Elsevier, respectively.
Carbon 07 00002 g005
Figure 6. Reaction time vs. quantum yields of CDs synthesized from CA by (a,c) hydrothermal method and (b,d) thermal decomposition. The data in (ad) are organized according to results in Ref. [57], Ref. [25], Ref. [30], and Ref. [58], respectively.
Figure 6. Reaction time vs. quantum yields of CDs synthesized from CA by (a,c) hydrothermal method and (b,d) thermal decomposition. The data in (ad) are organized according to results in Ref. [57], Ref. [25], Ref. [30], and Ref. [58], respectively.
Carbon 07 00002 g006
Figure 7. Reaction temperature vs. maximum emission of CDs synthesized from CA by (a) hydrothermal and (b) solvothermal methods. The data in (a,b) are organized according to Ref. [6] and Ref. [63], respectively.
Figure 7. Reaction temperature vs. maximum emission of CDs synthesized from CA by (a) hydrothermal and (b) solvothermal methods. The data in (a,b) are organized according to Ref. [6] and Ref. [63], respectively.
Carbon 07 00002 g007
Figure 8. The quantum yields of CDs synthesized from CA and different nitrogen sources. The data in (a,b) are organized according to Ref. [73] and Ref. [74], respectively.
Figure 8. The quantum yields of CDs synthesized from CA and different nitrogen sources. The data in (a,b) are organized according to Ref. [73] and Ref. [74], respectively.
Carbon 07 00002 g008
Figure 9. The photophysical properties of blue-, green-, and red-emitting of CDs in Ref. [63]. (ac) UV-visible absorption, excitation, and emission, (d) Raman spectra. Reproduced from Ref. [63] with permission of Copyright (2018) John Wiley and Sons.
Figure 9. The photophysical properties of blue-, green-, and red-emitting of CDs in Ref. [63]. (ac) UV-visible absorption, excitation, and emission, (d) Raman spectra. Reproduced from Ref. [63] with permission of Copyright (2018) John Wiley and Sons.
Carbon 07 00002 g009
Figure 10. CDs’ optical properties and potential applications.
Figure 10. CDs’ optical properties and potential applications.
Carbon 07 00002 g010
Table 1. The quantum yields of some CA-based CDs with different nitrogen (N) sources. QY: quantum yield.
Table 1. The quantum yields of some CA-based CDs with different nitrogen (N) sources. QY: quantum yield.
MethodsNitrogen SourcesEmissionQYsRef.
Hydrothermal method (200 °C, 8 h)Phenylalanine413 nm65%[20]
Microwave treatment (550 W, 7 min, aqueous solution)-440 nm7.2%[36]
Microwave treatment (700 W, 5 min, aqueous solution)Ethylenediamine455 nm41.3%[60]
Microwave treatment (700 W, 5 min, aqueous solution)Ammonia450 nm44.3%[37]
Hydrothermal method (200 °C, 5 h)-451 nm7.2%[6]
Hydrothermal method (180 °C, 4 h)Urea440 nm42.2%[62]
Hydrothermal method (200 °C, 5 h)Ethylenediamine470 nm60%[68]
Hydrothermal method (180 °C, 3 h)Dicyandiamide452 nm36.5%[69]
Hydrothermal method (200 °C, 4 h)L-cysteine432 nm75%[70]
Hydrothermal method (200 °C, 6 h)Tri(hydroxymethyl)
aminomethane
408 nm75%[71]
Hydrothermal method (180 °C, 4 h)Thiourea448 nm73.1%[72]
Table 2. The preparation and optical properties of some CA-based CDs. DMF: dimethylformamide.
Table 2. The preparation and optical properties of some CA-based CDs. DMF: dimethylformamide.
MethodsPrecursors
(Including CA)
EmissionQYsRef.
Microwave treatment (700 W, 40 s, aqueous solution)L-cysteine425 nm78%[77]
Hydrothermal method (180 °C, 8 h)Diethylenetriamine433 nm98%[78]
Hydrothermal method (200 °C, 5 h)Ethylenediamine445 nm80.6%[6]
Microwave treatment (400 W, 80 min, 150 °C, aqueous solution)Ethylenediamine445 nm82.99%[34]
Hydrothermal method (160 °C, 4 h)Ethylenediamine450 nm94%[30]
Microwave treatment (800 W, 4 min, aqueous solution)L-cysteine, dextrin495 nm22%[79]
Hydrothermal method (190 °C, 2 h)Urea455 nm29%[27]
510 nm30%
Thermal decomposition (170 °C, 70 min)Dicyandiamide528 nm73.12%[80]
Hydrothermal method (200 °C, 4 h)Ethylenediamine530 nm63.9%[65]
Hydrothermal method (200 °C, 12 h)Urea, ZnCl2580 nm51.2%[81]
Solvothermal method (180 °C, 4 h, formamide solution)Ethanediamine627 nm53%[75]
Microwave treatment (400 W, formamide solution)-640 nm22.9%[61]
Solvothermal method (160 °C, 6 h, DMF solution)Urea760 nm10%[76]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ren, J.; Malfatti, L.; Innocenzi, P. Citric Acid Derived Carbon Dots, the Challenge of Understanding the Synthesis-Structure Relationship. C 2021, 7, 2. https://doi.org/10.3390/c7010002

AMA Style

Ren J, Malfatti L, Innocenzi P. Citric Acid Derived Carbon Dots, the Challenge of Understanding the Synthesis-Structure Relationship. C. 2021; 7(1):2. https://doi.org/10.3390/c7010002

Chicago/Turabian Style

Ren, Junkai, Luca Malfatti, and Plinio Innocenzi. 2021. "Citric Acid Derived Carbon Dots, the Challenge of Understanding the Synthesis-Structure Relationship" C 7, no. 1: 2. https://doi.org/10.3390/c7010002

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