Haowen Sun,
Xirong Zhang,
Baojuan Wang,
Tianle He and
Huanming Xiong
*
Department of Chemistry, Shanghai Key Laboratory of Electrochemical and Thermochemical Conversion for Resources Recycling, Fudan University, Shanghai 200438, P. R. China. E-mail: hmxiong@fudan.edu.cn
First published on 14th October 2025
Porous materials (PMs) are a class of special materials characterized by their internal pore network structures, which afford them a larger specific surface area and a greater number of accessible active sites. Consequently, they find widespread applications in catalysis, energy storage and conversion, etc. Carbon dots (CDs), an emerging class of zero-dimensional carbon nanomaterials, possess highly cross-linked or carbonized cores along with abundant surface functional groups. The precursors of CDs are diverse and cost-effective, allowing for the tailoring of specific structures through controlled reaction conditions. Initially, CDs were extensively utilized in bioimaging and fluorescence detection due to their characteristic photoluminescence properties. However, in recent years, a substantial body of research has focused on employing CDs as fundamental building blocks or modifying species to fabricate various PMs, with experimental evidence underscoring their significant role. In this review, PMs are categorized into porous carbons, porous inorganic materials, and porous gel materials based on their fundamental constituents. We summarize recent advances in PMs constructed using CDs, with a particular emphasis on the influence of CDs regarding the morphology and pore structure of these materials, as well as the underlying mechanisms. This systematic overview aims to provide new insights into the design of porous materials and the multifunctional applications of CDs.
Carbon dots (CDs) are an emerging class of fluorescent carbon nanomaterials characterized by their ultrasmall size, typically not exceeding 10 nm.15,16 Most studies agree that the structure of CDs consists of two components: a highly cross-linked or carbonized core and a shell enriched with surface functional groups and defects.17 Compared to the core, the surface states exert a more profound influence on the properties of CDs, particularly their photoluminescence behavior.18–20 CDs are generally synthesized from precursors containing conjugated structures or readily cross-linkable organic molecules. The abundance and low cost of such precursors allow for flexible modulation of the CD structure.21–23 Owing to their unique characteristics, CDs have been extensively employed as fundamental building blocks or functional modifiers in the construction of various materials, exerting significant influence on the morphology and pore structures of the resulting products.
While previous reviews have summarized composite strategies and host–guest interactions between CDs and porous materials,24,25 it is increasingly recognized that CDs also play a crucial role in pore formation and morphology control. Therefore, this review specifically emphasizes the role of CDs in regulating material morphology, pore structures, and SSA. It covers three main aspects: (i) porous carbon materials engineered by CDs; (ii) porous inorganic materials constructed using CDs; and (iii) porous polymer materials induced by CDs (Fig. 1). In the end, we provide a summary and discuss the primary challenges and future opportunities for CDs in building innovative materials.
As illustrated in Fig. 2a, after nearly two decades of development, CDs are generally classified into four types: graphene quantum dots (GQDs), carbon nanodots (CNDs), carbon quantum dots (CQDs), and carbonized polymer dots (CPDs), which are generally regarded as possessing a core–shell structure.30–33 GQDs are made up of single layers of graphene, which exhibit a high degree of graphitization, extensive sp2 domains, and lattice spacings ranging from 0.18 to 0.24 nm—values that are comparable to those of bulk graphite.34 In contrast, CQDs are quasi-spherical carbon nanoparticles which exhibit a crystalline core based on a mixture of sp2 and sp3 carbons.35 Compared to GQDs, CQDs have a greater abundance of surface functional groups and higher defect densities. Due to their high crystallinity, the optical bandgaps of both GQDs and CQDs are significantly influenced by their size.36 Due to their high crystallinity, GQDs and CQDs are often prepared via top-down approaches, such as chemical oxidation or arc discharge, which exfoliate them from bulk carbon materials. CNDs are often confused with CQDs. However, some researchers distinguish CNDs as quasi-spherical carbon nanoparticles characterized by an amorphous core, as opposed to the more crystalline core found in CQDs.37,38 In recent years, an increasing number of studies have adopted bottom-up approaches to synthesize CDs, whereby organic small molecules undergo polymerization and cross-linking to form distinctive core–shell structures (Fig. 2b).39–41 The majority of CDs prepared via this method are classified as CPDs, a term proposed by Yang et al. in 2018.42 Due to incomplete carbonization, CPDs exhibit low crystallinity and possess typical polymeric characteristics.43 CPDs are characterized by the highest abundance of surface functional groups among CD types, enabling facile functionalization or interaction with other materials through tailored precursor selection and reaction conditions.19,44,45
![]() | ||
| Fig. 2 (a) Four categories of CDs and their respective characteristics.49 Reproduced with permission from ref. 49. Copyright 2022 Springer Nature. (b) Diagram to describe the reaction process of hydrothermal cross-linking polymerization to prepare CPDs by the “bottom-up” route. Reprinted with permission from ref. 39. Copyright 2019 American Chemical Society. | ||
Owing to their unique array of properties, CDs have been increasingly utilized by researchers as fundamental building blocks or additives in material synthesis. Experimental results demonstrate that CDs exert a significant influence on the morphological features of materials, particularly their pore structures.46–48 (Unless otherwise specified, the term “CDs” in the subsequent discussion refers specifically to CPDs.)
In 2015, Hou et al. developed a novel strategy for the large-scale synthesis of carbon dots via an aldol condensation reaction.50 Without any purification (retaining residual NaOH), the as-prepared products were annealed at 800 °C for 2 hours under an argon atmosphere. Under such high temperature and Na catalysis effects, the solid products were rapidly decomposed to produce abundant carbon atoms which then gradually self-assembled to form carbon nanosheets. The rapid decomposition of abundant surface functional groups on CDs generates transient high internal pressure, which effectively suppresses the stacking of carbon nanosheets during their self-assembly process. Finally, with the linkage of numerous O-rich functional groups, the as-grown nanosheets exhibited a cross-linked 3D framework structure (Fig. 3a and b). In subsequent work, the researchers purified as prepared CDs and pyrolyzed them with sodium dihydrogen phosphate (NaH2PO4), yielding phosphorus-doped, large-area porous carbon nanosheets.51 It confirmed that the method possesses universality and allows for facile heteroatom doping. The resulting anode material for sodium-ion batteries (SIBs) delivered a remarkable specific capacity of 108 mAh g−1 even at 20 A g−1, attributed to its abundant micropores (1.2 nm) and expanded interlayer spacing.
![]() | ||
Fig. 3 (a) SEM and (b) nitrogen adsorption–desorption isotherms of 3D PCFs. Reproduced with permission from ref. 50. Copyright 2015, Wiley-VCH. (c) General synthetic route of the ultra microporous carbons. Reprinted with permission from ref. 52. Copyright 2018, Wiley-VCH. SEM images of the samples at different reactant ratios (NaOH : acetylacetone): (d) and (g) 1 : 10, (e) and (h) 3 : 10, (f) and (i) 5 : 10. Reprinted with permission from ref. 54. Copyright 2020, Elsevier. (j) Schematic illustrations of the ice templated assembly strategy for fabricating 3D porous carbon. Reprinted with permission from ref. 56. Copyright 2023, Elsevier. | ||
To emphasize the advantages of CDs over conventional carbon sources as precursors for PMs, Zhang et al. compared and analysed the structure and properties of PCMs derived from coal and coal-derived GQDs.52 As illustrated in Fig. 3c, the coal-derived GQDs were prepared via chemical oxidation. Benefiting from their ultrasmall size, minimal amounts of KOH could be uniformly dispersed at the edges or interlayers of GQDs. Subsequent high-temperature treatment enabled in situ activation of GQDs by KOH from the interior to the exterior, yielding ultra microporous carbon (CoDCs) with both a high SSA (1730 m2 g−1) and an exceptional packing density (0.97 g cm−3). In contrast, direct use of coal as the carbon precursor not only consumed significantly more KOH (eightfold higher than that for GQDs) but also resulted in PCMs with heterogeneous pore size distribution and a markedly lower packing density (0.4 g cm−3). When CoDCs were employed as electrode materials, the optimized PCMs achieved an ultrahigh areal capacitance of 5.70 F cm−2 at 1 A g−1 under a high mass loading of 25 mg cm−2. Similarly, Huang et al. systematically compared sulfur-doped hard carbon derived from CDs (2SC-600), glucose, and carbon fibers.53 The results demonstrated that CDs not only facilitated pore generation but, more critically, provided abundant sulfur-binding sites on their surfaces. The formed C–S bonds decomposed into H2S, generating more micropores. Benefiting from the selective substitution of oxygen atoms in the carbon framework by sulfur, 2SC-600 exhibits a high initial coulombic efficiency of 68%.
Beyond their role as chemical porogens, alkalis can serve as crystalline templates. As demonstrated by Qiu et al., hollow carbon microboxes were synthesized through precise control of NaOH concentration and carbonization temperature.54 Under critical thermal conditions, NaOH underwent transformation into Na2CO3 crystalline templates (eqn (1)), which directed the assembly of carbon layers into polyhedral architectures via synergistic electrostatic and van der Waals interactions. The template-precursor balance proved essential: insufficient NaOH led to incomplete template formation, while excess NaOH generated oversized crystals that prevented uniform carbon coating on polyhedral units, ultimately causing structural fragmentation into sharp-edged carbon nanosheets (Fig. 3d–i).
| 6NaOH + 2C → 2Na + 3H2 + 2Na2CO3 | (1) |
Notably, benefiting from abundant heteroatom doping, CDs can form PCMs without porogens. In 2014, Zhou et al. fabricated mesoporous carbons (MCs) with a high surface area (183.6 m2 g−1) by simply drop-casting CQD solutions onto silicon wafers followed by direct pyrolysis.55 Remarkably, MCs demonstrated outstanding catalytic selectivity (84.67%) in the oxidation of cyclooctene to epoxy cyclooctane with 17.37% conversion. As a clearer example, Shi et al. employed a mild chemical oxidation method to prepare CQDs from coal tar pitch.56 Subsequently, ice-template induced assembly coupled with the carbonization strategy reveals a transition from 0D CQDs to 3D porous carbon frameworks (PCF-800) composed of cross-linked and twisted nanosheets (Fig. 3j). Through comparative analysis of different pyrolysis stages, the authors proposed that the thermal decomposition of surface functional groups on CQDs and small molecules generates trace gases at high temperatures, thereby creating additional pores in the carbon skeleton.
The sol–gel method is a commonly employed approach for synthesizing PCMs. However, during pyrolysis, polymer chains are converted into a carbonaceous framework, often leading to issues such as pore collapse or structural inhomogeneity. Fortunately, these challenges can be mitigated by introducing CDs, which act as pore-directing agents and stabilize the porous network during carbonization. For instance, Zhang et al. synthesized Co9S8/CD@NSC by oxidizing a mixture of aniline, CDs, and CoCl2·6H2O to form a polyaniline hydrogel, followed by freeze-drying and dual-step annealing.57 The resulting material exhibited a high SSA of 745.2 m2 g−1, with both CDs and CoCl2 contributing to pore formation. Notably, the SSA increased with the amount of CDs added (Fig. 4a and b), attributed to the crosslinking between CDs and polyaniline nanofibers, as well as the role of CD surface functional groups in dispersing CoCl2. Similarly, Xu et al. introduced CDs into a starch-based hydrogel (CDSA), where the oxygen-containing functional groups on the CDs formed hydrogen bonds with hydroxyl groups on starch chains, promoting polymerization and crosslinking.58 As illustrated in Fig. 4c–h, CDs acted as structural stabilizers, preventing pore collapse during both freeze-drying and carbonization. As a result, PCMs derived from CDSA (C-CDSA) exhibited a higher SSA and a more developed porous structure compared to the CD-free counterpart (C-SA).
![]() | ||
| Fig. 4 (a) Nitrogen adsorption–desorption isotherms and (b) pore size distribution of Co9S8/CD@NSC with different amounts of CD addition. Reprinted with permission from ref. 57. Copyright 2019, American Chemical Society. Photographs and SEM images of the cross-section of (c)–(e) C-SA and (f)–(h) C-CDSA. Reprinted with permission from ref. 58. Copyright 2022, Royal Society of Chemistry. (i) Nitrogen adsorption–desorption isotherms of PD/GS composites and GS. FESEM of (j) GS and (k) PDs/GS-0.5. Reprinted with permission from ref. 60. Copyright 2017, Elsevier. The TEM image of the (l) N,P-PCNFs and (m) PCNFs. Reprinted with permission from ref. 62. Copyright 2024, Royal Society of Chemistry. (n) Nitrogen adsorption/desorption isotherms; (o) pore size distributions of CDs@rGO composites treated with HI vapor for various reduction times. Reprinted with permission from ref. 64. Copyright 2019, Elsevier. | ||
Graphene is a remarkable material composed of carbon atoms arranged in a two-dimensional honeycomb lattice through sp2 hybridization. This unique structure grants it exceptional optical, electrical, and mechanical properties, positioning it as a revolutionary material for future technologies.59 However, a significant challenge arises from its tendency to restack due to strong π–π interactions between its sheets, which reduces SSA and limits its practical applications. Fortunately, as 0D carbon materials, CDs can readily integrate with graphene, effectively mitigating this issue. Wei et al. employed a sol–gel approach to fabricate composites of CDs and graphene nanosheets (PDs/GS-0.5).60 The incorporation of CDs effectively transformed the initially smooth and planar graphene into fractured and wrinkled morphologies, accompanied by a significant increase in SSA (Fig. 4i–k). Meanwhile, this morphology introduces various microporous structures, which is highly beneficial for enhancing the specific capacitance. Heteroatom doping of graphene is a common strategy to enhance its properties. Zhang et al. employed N,S co-doped CDs (N,S-CDs) in a hydrothermal reaction with reduced graphene oxide, achieving not only effective heteroatom incorporation but also inducing the assembly of a morphology with a significantly enlarged SSA.61
In recent years, porous carbon nanofibers (PCNFs) have emerged as novel 1D carbon materials, which can be synthesized via methods such as chemical vapor deposition or electrospinning. PCNFs not only inherit the excellent electrical conductivity and mechanical flexibility characteristic of 1D nanostructures, but their porous features further expand their potential applications. To achieve this goal, Yuan et al. introduced polytetrafluoroethylene and N,P-co-doped CDs (N,P-CDs) into the precursor solution used for electrospinning.62 The resulting fibrous membrane was annealed at 900 °C for 2 hours, yielding N,P-PCNFs with a distinct bubble-like porous morphology. Compared to PCNFs prepared without the addition of N,P-CDs, the N,P-PCNFs exhibited uniform porosity and no observable fiber breakage, indicating that the incorporation of N,P-CDs significantly enhanced the structural stability of CNFs (Fig. 4l and m). In 2025, Qin et al. employed a similar approach to fabricate CoOx-loaded porous CNFs (CNF@CDs@CoOx), in which CDs also played a crucial role in enhancing the mechanical flexibility of the fibers.63 When used as an anode material for SIBs, the resulting CNF@CDs@CoOx demonstrated promising potential for application in flexible energy storage devices.
In some circumstances, the insufficient surface functional groups on CDs fail to contribute to the formation or modulation of the pore structure. However, CDs possess a highly stable structure, enabling them to serve as “nano-buttresses” that physically reinforce the carbon framework. Hoang et al. synthesized CDs@rGO composites via a hydrothermal method and found that during the HI-mediated reduction of GO, CDs served as nanofillers to effectively inhibit the restacking of graphene sheets, thereby preserving the porous architecture to a significant extent (Fig. 4n and o).64 Metal–organic frameworks (MOFs), with their intrinsic porosity and highly tunable elemental composition, are considered ideal precursors for porous carbon materials. However, under extreme temperatures, the framework structure of MOFs tends to collapse, leading to pore blockage. Tang et al. reported PCMs derived from MOF-5, in which GQDs with abundant carboxyl groups and rigid structure could uniformly distribute in the MOF-5 precursor by coordinating with [Zn4O]6+ clusters and effectively reinforce the carbon skeleton during pyrolysis (Fig. 5a).65 As a result, the SSA of the obtained PCMs was 500 m2 g−1 higher than that of PCMs derived from pristine MOF-5.
![]() | ||
| Fig. 5 (a) The preparation and structural transformation process of MPC and GMPC-0.35. Reprinted with permission from ref. 65. Copyright 2022, Elsevier. (b) Pore size distribution of PNDCs-500CDs-y °C with different carbonization temperatures. (c) TEM of PNDCs-500CDs-1200 °C. (d) SAXS patterns of PNDCs, PNDCs-500CDs, and PNDCs-500CDs-1200 °C. Reprinted with permission from ref. 66. Copyright 2024, Wiley-VCH. (e) Schematic representation of the synthesis of hollow N-doped carbon. (f) and (g) TEM and HRTEM images of p-HNCs. Reprinted with permission from ref. 67. Copyright 2019, Elsevier. | ||
Closed pores represent a distinct type of pore structure compared to open pores. Although they are isolated from the external environment, they play a crucial role in anode material SIBs. Huang et al. synthesized composite nanosheets of polyimide and CDs (PI-xCDs) via a solvothermal method.66 During subsequent high-temperature carbonization, the surface functional groups on the CDs decomposed and released gases, leading to the formation of flower-like porous carbon nanosheets (PNDCs-xCDs-y °C). Interestingly, although BET measurements failed to detect porosity when the carbonization temperature exceeded 1000 °C, TEM observations revealed the presence of abundant mesopores (Fig. 5b and c). Furthermore, small-angle X-ray scattering (SAXS) analysis of PNDCs-500CDs-1200 °C displayed distinct peaks in Q = 0.1–1 Å−1, indicating a large number of closed pores (Fig. 5d). These findings suggest a transformation from micropores to closed pores with increasing carbonization temperature. The presence of such closed-pore structures provides additional sodium storage sites in the plateau region for the high storage capacity.
Hong et al. dispersed CDs in a water/ethanol mixed solvent to form micelles, which were then coated with a polypyrrole shell via chemical oxidative polymerization.67 As shown in Fig. 5e, the abundant functional groups on the CDs provide ample reactive sites for the polymerization of pyrrole, leading to the formation of a uniform shell layer. Subsequently, CDs were removed by repeated washing with ethanol and water, followed by annealing at 650 °C for 5 h to obtain w-HNCs. Samples in which CDs were not removed were denoted as p-HNCs. Both p-HNCs and w-HNCs possess hollow structures (Fig. 5f and g), but p-HNCs exhibit more micropores, which originates from the internal pressure and gas release generated during the decomposition of CDs. The enhanced porosity was advantageous for electrolyte infiltration and ion adsorption. As a result, anode materials constructed using p-HNCs for potassium-ion batteries delivered high reversible specific capacities of 254 mAh g−1 at 0.1 A g−1 after 100 cycles and 160 mAh g−1 at 1.0 A g−1 after 800 cycles.
For partially carbonized CDs, a substantial number of active chemical bonds remain within their structure. These residual bonds can serve as targets for directional etching, enabling precise modulation of pore size and uniformity. Huang et al. demonstrated that CDs can serve as hard templates for the fabrication of PCMs with highly uniform pore sizes.68 In this work, coal tar-derived CPDs were reacted with melamine to form a hyper-crosslinked polymer (HCP). Acting as sacrificial templates, the CPDs underwent targeted “excavation” by KOH, producing pores that closely matched CPD dimensions (Fig. 6a). As a result, the carbonized product exhibited a highly developed porous structure dominated by small mesopores (2–10 nm accounting for 92.07%) and an ultrahigh SSA of 4241 m2 g−1. In 2024, Zhang et al. reported the synthesis of sub-nanometer PCMs using CDs prepared from citric acid and diethylenetriamine (DETA), which contained a large number of amide bonds in its interior structure.46 These CDs were crosslinked into a gel via reaction between surface amine groups and poly(ethylene glycol) diglycidyl ether (PEGDGE), followed by immersion in KOH. As shown in Fig. 6b, the internal amide bonds within the CDs were selectively cleaved, leading to their fragmentation into small pieces. After carbonization, the resulting PCMs exhibited numerous sub-nanopores (0.64–0.8 nm), which are favourable for K+ desolvation and rapid ion transport (Fig. 6c). As a result, the optimal sample with a high packing density (0.81 g cm−3) displays outstanding capacitances (gravimetric 515.5 F g−1, areal 5.16 F cm−2, and volumetric 417.6 F cm−3 respectively at 1 A g−1) at the commercial-level mass-loading of 10 mg cm−2.
![]() | ||
| Fig. 6 (a) Schematic diagram showing the transformation of CCPD@HCPM(CC). Reprinted with permission from ref. 68. Copyright 2023, Elsevier. (b) Schematic diagram of the synthetic procedure for CPDHs. (c) Schematics of K+ residing in pores with different sizes. Reprinted with permission from ref. 46. Copyright 2025, Wiley-VCH. | ||
![]() | ||
| Fig. 7 (a) TEM images of S-C3N4/C-dot. (b) Pore size distribution of S-C3N4/C-dot with addition of different C-dots. Reprinted with permission from ref. 70. Copyright 2019, Elsevier. (c) and (d) High-resolution SEM image of CDs@P-Eu-MNs. Reprinted with permission from ref. 71. Copyright 2022, Elsevier. (e) The role of CDs in modified Mxene. Reprinted with permission from ref. 47. Copyright 2023, Springer Nature. (f) Different morphologies resulting from different concentrations of CDs (From left to right, the concentrations of CDs are respectively 0, 0.25, 0.5, and 1 mg mL−1). Reprinted with permission from ref. 72. Copyright 2016, Wiley-VCH. | ||
Wang et al. synthesized CDs via a hydrothermal reaction using citric acid and cysteine, and subsequently reacted them with citric acid and Eu(NO3)3·6H2O to fabricate CD-based porous europium micro-networks (CDs@P-Eu-MNs).71 As shown in Fig. 7c and d, CDs not only serve as excellent nucleation and growth platforms for the initial nanostructures, but also contribute to the generation of abundant pores through dehydration reactions between their surface functional groups and Eu-MNs. The construction of CDs@P-Eu-MNs not only extends light absorption into the visible region but also enhances the separation and transport of photogenerated charge carriers. Similarly, Wang et al. introduced CDs into layered Ti3CNTx MXene via simple stirring, followed by annealing at 350 °C for 1h to obtain porous MXene.47 As illustrated in Fig. 7e, on one hand, CDs act as interlayer spacers to prevent restacking of MXene nanosheets, and on the other, CDs partially decompose to release NH3 which in situ etches the MXene to form a 3D interconnected porous network. The resulting flexible film electrodes achieve excellent gravimetric capacitance (688.9 F g−1 at 2 A g−1) and outstanding rate capability.
As early as 2016, Wei et al. utilized CDs to modulate the morphology of NiCo2O4 grown on Ni foam.72 By increasing the concentration of CDs in the precursor solution from 0 mg mL−1 to 1 mg mL−1, the NiCo2O4 morphology evolved from sea urchin-like to flower-like, and eventually to bayberry-like structures, accompanied by significant changes in SSA and the pore structure (Fig. 7f). By simply tuning the reaction conditions, PIMs with diverse morphologies can be readily obtained, which has attracted significant attention from researchers. In 2017, Wei et al. introduced CQDs into MxOy (M = Co, Ni) precursor solutions for hydrothermal synthesis, followed by annealing the resulting powders at 300 °C for 3 h.73 Compared to MxOy, CQDs/MxOy exhibit distinctly different morphologies. It was found that CQDs could induce the self-assembly of MxOy nanosheets or nanowires into flower-like or sea urchin-like architectures (Fig. 8a–f) through electrostatic interactions with metal ions, leading to a substantial increase in both SSA and specific capacitance.
![]() | ||
| Fig. 8 FESEM images of MxOy: (a) Co3O4, (c) NiCo2O4, (e) NiO and the CQDs/MxOy: (b) CQDs/Co3O4, (d) CQDs/NiCo2O4, (f) CQDs/NiO. Reprinted with permission from ref. 73. Copyright 2017, Elsevier. TEM images of CT-500 products synthesized at different reaction time: (g) 0.5 h, and (h) 1.5 h of hydrolysis reaction at room temperature, (i) 1 h, (j) 3 h, (k) 6 h, and (l) 12 h of hydrothermal time at 180 °C. Reprinted with permission from ref. 74. Copyright 2020, Elsevier. (m) Schematic illustration of Bi2Se3 self-assembly induced by CDs. Reprinted with permission from ref. 76. Copyright 2022, Elsevier. | ||
Although the modulation of inorganic materials by CDs is a commonly observed phenomenon, a more detailed investigation into their morphological evolution is still required. In 2020, Wu et al. reported the use of CQDs to regulate the growth of TiO2.74 Specifically, TiO2 nanorods were mixed with CQDs under stirring, followed by hydrothermal treatment, resulting in flower-like architectures rich in mesopores. The morphological evolution process is shown in Fig. 8g–l: after CQDs incorporation, the originally dense TiO2 nanorods became loose and uneven (Fig. 8h). Following just 1 hour of hydrothermal treatment, the morphology rapidly transformed into oval-shaped structures with burr-like surfaces. The results suggest that the incorporation of CQDs can decrease the surface energy of the system, then promoting the formation of mesoporous TiO2 nanostructure. With further increasing the hydrothermal reaction time, the product showed a spherical morphology, where more nanosheets grew onto the surface of nanospheres. The obtained CT-500 exhibited a SSA of 197.7 m2 g−1 along with abundant mesopores. Song et al. has also synthesized TiO2 spheres with flower-like surfaces via a solvothermal reaction between nitrogen-doped CDs (N-CDs) and TiO2.75 They found that the amount of NCDs plays a critical role in determining the final morphology. The optimized TiO2/NCDs5 sample, featuring ideal morphology and pore size distribution, was employed as a protective layer for zinc-ion battery anodes, which facilitated ion transport, suppressed dendrite growth, and significantly improved cycling stability. Wang et al. synthesized uniform rod-like Bi2Se3/CDs composites using ZnSe(DETA)0.5 microbelts as a substrate via an ion-exchange method.76 Through detailed morphological evolution studies, the authors proposed a CDs-induced self-assembly growth mechanism (Fig. 8m): when sufficient CDs are incorporated in the reaction, firstly, they will coat on the ZnSe(DETA)0.5 microbelts due to the coordination effects. After adding BiCl3, the CDs will interact with Bi3+ by electrostatic interaction, which drives the CDs to act as active sites for ion exchange process between Bi3+ and Zn2+, yielding the formation of Bi2Se3 nanosheets tightly bound with CDs and distributed along the nanobelts. The crosslinked Bi2Se3 nanosheets form a robust 3D framework for fast-kinetics lithium storage, providing efficient pathways for electron and ion transport.
Notably, beyond guiding the assembly of inorganic materials, CDs impart additional functions. For instance, Liu et al. also prepared SnS2/CDs composites via hydrothermal synthesis, which were applied as fillers in polymer electrolytes for lithium-ion batteries (LIBs).77 The incorporation of CDs not only induced the transformation of SnS2 from a stacked hexagonal structure into a flower-like morphology but also promoted the generation of abundant sulfur vacancies, which significantly facilitate dissociation of lithium salt and accelerate migration of Li+. Furthermore, CDs dotted on the surface of SnS2 have rich organic functional groups, which can serve as the bridging agent to enhance the compatibility of filler and polymer, leading to superior mechanical performance and fast ion transport pathway.
In summary, CDs have been demonstrated to effectively direct the growth of inorganic materials, preventing their excessive aggregation and thereby generating higher SSA and well-defined pores. Moreover, CDs can also be simply composited with inorganic materials followed by calcination, where their nanoscale dispersibility and the thermal decomposition of surface functional groups contribute to the formation of uniform pores.
![]() | ||
| Fig. 9 (a)–(e) SEM images of hydrogels: 1CDs@EW, 4CDs@EW, CEWH (8CDs@EW), and 12CDs@EW (the numbers represent the concentration of CDs added.). Scale bars: 20 μm. (f) A proposed mechanism for the formation of a stretchable and transparent network structure using CDs and EW peptide chains. Reprinted with permission from ref. 48. Copyright 2023, American Association for the Advancement of Science. SEM images of (g1) and (g2) TAC0 without ACDs and ACDs-loaded nanocomposite hydrogels: (g3) and (g4) TAC100 with 100, (g5) and (g6) TAC500, (g7) and (g8) TAC1000 (the numbers indicate the amount of ACDs added, in μL). Reprinted with permission from ref. 79. Copyright 2022, American Chemical Society. (h) The schematic comparison between the GPEs derived from three polymers: the cast PVDF-HFP (solid polymer), the phase inversion derived PVDF-HFP (stacked polymer dots), and the PVDF-HFP/CDs (polymer-CDs columns). Reprinted with permission from ref. 103. Copyright 2024, Elsevier. | ||
In fact, some functional groups on the CDs do not participate in crosslinking with the monomers. Instead, they may interact with the side groups of the polymer chains, leading to a reduction in the original pore size of PGMs. However, this effect is not necessarily detrimental. Das et al. synthesized Apium graveolens-derived carbon dots (ACDs) and subsequently prepared thermoplastic starch/poly(acrylic acid)/ACD hydrogels via thermal polymerization.79 As shown in Fig. 9g1–g8, with the gradual increase of ACD addition from 0 μL to 1000 μL, the pore size of the hydrogel progressively decreased. This effect is attributed to the formation of additional physical crosslinks through hydrogen bonding between ACDs and the polymer chains. Moreover, the room-temperature self-healing behaviour was observed for the ACDs-reinforced hydrogels, with a healing efficacy of more than 45%. Similarly, Alzahrani prepared CDs via a thermal coupling reaction between polyethyleneimine and L-cysteine, followed by their incorporation into a precursor solution to fabricate the chitosan/poly(acrylamide-co-2-aminoethyl methacrylate) hydrogels.80 The introduction of CDs led to a reduction in pore size and a denser gel network. This phenomenon can be explained by the formation of broad secondary interactions between the polarities on the surface of CDs and the additional groups attached to the hydrogel polymer chains. The resulting hydrogel not only inherits the antibacterial and antioxidant properties of CDs but also effectively prevents the penetration of moisture and gases, demonstrating excellent potential in food preservation applications.
To provide readers with a clear and systematic overview, we have summarized in Table 1 the key information and application of CD-fabricated PMs.
| Sort | Material | Precursors of CDs | Fabrication methods of PMs | SSA (m2 g−1) | Application | Ref. |
|---|---|---|---|---|---|---|
| PCMs | MCs | Graphite | Direct calcination of CDs | 183.6 | Selective oxidation of hydrocarbons | 55 |
| 3D PCFs | Acetone | Direct calcination of CDs | 467 | Sodium-ion batteries (SIBs) | 50 | |
| PDs/GS | Polyethyleneimine (PEI) | Ultrasonic compounding of CDs with graphene nanosheets | 407.2 | Supercapacitor | 60 | |
| 1D CNF, 2D CNS, and 3D CFW | Acetone | Calcination of CDs and salt mixtures | 429.5 | SIBs | 81 | |
| NPOCD/HPC | Phytic acid, PEI | Calcination of CDs@Polyacrylamide hydrogel | 1025 | Supercapacitor | 82 | |
| PS–BC | Benzenesulfonic acid, phenylphosphonic acid, acetone | Direct calcination of CDs | 671.6 | Na- and Li-ion batteries (LIBs) | 83 | |
| p-HNCs | Acetaldehyde | Calcination of CDs-assisted hollow polypyrrole | 68.78 | Na- and K-ion batteries (KIBs) | 67 | |
| 3@10–600 °C | Acetylacetone | Na2CO3-assisted templating method | 60.6 | SIBs | 54 | |
| CCPD@HCPM(CC) | Coal tar | Oriented etching of CDs by KOH | 4241 | Supercapacitor | 68 | |
| PNDCs-500CDs-1200 °C | Acetaldehyde | Calcination of CDs@polyimide | — | SIBs | 66 | |
| CPDHs | Citric acid (CA), diethylenetriamine | CDs gelation followed by calcination | 1020.4 | Supercapacitor | 46 | |
| N-PCFs | CA, urea | Direct calcination of CDs | 483.7 | LIBs | 84 | |
| CDs/NSPC | Bovine serum albumin | Direct calcination of CDs | 1172.54 | Zn–air batteries(ZABs) | 85 | |
| PC-CDs | P-nitroaniline | Ultrasonic compounding of CDs with p-nitroaniline derived PCMs | 2462 | Supercapacitor | 86 | |
| DAEC | Yolks | Calcination of CDs and egg white mixtures | 1999.3 | Supercapacitor | 87 | |
| GMPC-0.35 | Coal powder | CDs encapsulated in MOF-5 followed by calcination | 1841 | Supercapacitor | 65 | |
| PCF-700 | Coal tar pitch | Direct calcination of CDs | 65.21 | LIBs | 56 | |
| CQDs/HPC | CA, urea | Calcination of CDs with camellia oleifera shell powder and urea | 1750 | Supercapacitor | 88 | |
| ACD | Glucose | Direct calcination of CDs | 1246 | Supercapacitor | 89 | |
| rGO/CDs | CA | Ultrasonic-assisted blending of CDs with rGO | 44.52 | Supercapacitor | 90 | |
| PIMs | S-C3N4/C-dot | Glucose | Calcination of S-C3N4/C-dot composite | 98 | Photocatalysis | 70 |
| Ni9PCN | CA, urea | Calcination of CDs with urea | 47.603 | Photocatalysis | 91 | |
| CDs-MnO2 | Ammonium citrate, sodium dihydrogen phosphate | Microwave-assisted hydrothermal reaction of CDs and KMnO4 | 254.83 | Oxygen evolution reaction (OER) | 92 | |
| Ni(OH)2/af-GQDs | Graphite | Homogeneous blending of CDs with Ni(OH)2 nanosheets | 84.6 | Supercapacitor | 93 | |
| Black ZnO Nano Clusters | CA, ethylenediamine(EDA) | Reaction of CDs with ZnO precursors | 49.8 | Nickel-zinc alkaline batteries | 94 | |
| Bi2Se3/CDs | Acetaldehyde | CD-assisted hydrothermal growth of Bi2Se3 on ZnSe(DETA)0.5 substrates | 59.89 | LIBs | 76 | |
| Mn3O4/NCDs | CA, EDA | Hydrothermal reaction of CDs with Mn3O4 precursors followed by calcination | 45.5 | Zinc-ion batteries (ZIBs) | 95 | |
| SnO2/CDs-300 | Acetaldehyde | Solvothermal reaction of CDs with SnO2 precursors in ethanol | 77.61 | LIBs | 96 | |
| LDH/CD@CC | Cetylpyridinium chloride monohydrate | Electrodeposition using mixed electrolyte containing Ni(NO3)2, Co(NO3)2, and CDs | — | Supercapacitor | 97 | |
| Bi2O3/NCDs | CA, EDA | Solvothermal reaction of CDs with Bi(NO3)3·9H2O followed by calcination | — | Supercapacitor | 98 | |
| NCDs@g-C3N4 | p-phenylenediamine | Hydrothermal reaction of CDs with g-C3N4 | 60.9 | Photocatalysis | 99 | |
| PGMs | PLLA/CQD | Corn stalks | Gelation of CDs and poly(L-lactic acid) followed by solvent evaporation | — | Cell culture | 100 |
| FH-5 | CA, polyvinyl alcohol | Synthesis of carboxymethylated carbon nanofibers grafted with CDs and free radical copolymerization with acrylic acid monomer | — | Contaminant treatment | 101 | |
| PCCEs | CA, EDA | Polymerization of PVDF and PEG-CDs by solvent evaporation method | — | Gel electrolytes of LIBs | 102 | |
| CEWH | CA, urea | Thermal polymerization of CDs and egg white | — | Wound dressing | 48 | |
| PVDF-HFP/U-CDs | CA, urea | Polymerization of PVDF and CDs by a mild phase inversion process | — | Gel electrolytes of LIBs | 103 | |
| GCDs-PAA | Garlic | Thermal polymerization of CDs, acrylic acid, pectin and polyethylene glycol dimethacrylate | — | Wound dressing | 104 | |
| CD-HY | Chitosan, poly(ethyleneimine) | Gelation of chitosan–CD–glycerol mixture induced by NaOH | — | Cell monitoring | 105 | |
| CDs/HD | Tartaric acid, triethylenetetramine | Thermal polymerization of CDs, polyvinyl alcohol, polyvinylpyrrolidone and citric acid | — | Contaminant treatment | 106 |
(i) The true structure of CDs remains unclear, and existing structural models are primarily derived from indirect evidence obtained through various characterization techniques. A more accurate understanding of the structural features of CDs would facilitate deeper insights into their role in the synthesis of PMs and enable a clearer analysis of structure–property relationships. Employing structurally well-defined precursors in combination with advanced characterization methods may offer a promising strategy to address this challenge.
(ii) The large-scale and precise synthesis of CDs remains a critical challenge that directly impacts their potential for industrial application. Currently, solvothermal methods dominate CD synthesis in the laboratory, offering good particle uniformity. However, the high energy consumption and stringent safety requirements limit large scale production. In recent years, the salt-assisted solid-state pyrolysis method has emerged as a promising alternative, eliminating the hazards of high-pressure conditions and enabling kilogram-scale production.
(iii) Most of the research on PMs constructed using CDs is application-oriented, ignoring the underlying reaction mechanisms and structure–property relationships. Although some studies have investigated the intermediate products at various reaction stages, the overall synthesis process remains a “black box,” as ex situ characterization techniques frequently miss critical transient information. Advanced in situ characterization and real-time observation techniques may be helpful. A more comprehensive understanding of the mechanisms would facilitate the broader application of CDs across diverse materials.
(iv) Although the potential of CDs in constructing porous materials has been well established in recent years, only a few studies have investigated the influence of CDs on other features of PMs, such as the electronic structure, which is also critical for PM performance. Moreover, CDs possess unique properties, including fluorescence, catalytic activity, and bioactivity. Integrating these properties with PMs to create multifunctional materials is therefore highly meaningful.
Nevertheless, CDs have demonstrated excellent potential in constructing PMs. We believe that in the near future, these challenges will be conquered by researchers, and CDs will make greater contributions to the preparation and application of PMs.
D Network of Graphene-Oxide Nanosheets Decorated with Carbon Dots for High-Performance Supercapacitors, ChemSusChem, 2017, 10, 2626–2634 CrossRef CAS PubMed.| This journal is © the Partner Organisations 2025 |