Skip to main content

Green synthesis of glyco-CuInS2 QDs with visible/NIR dual emission for 3D multicellular tumor spheroid and in vivo imaging

Abstract

Glyco-quantum dots (glyco-QDs) have attracted significant interest in bioimaging applications, notably in cancer imaging, because they effectively combine the glycocluster effect with the exceptional optical properties of QDs. The key challenge now lies in how to eliminate the high heavy metal toxicity originating from traditional toxic Cd-based QDs for in vivo bioimaging. Herein, we report an eco-friendly pathway to prepare nontoxic Cd-free glyco-QDs in water by the “direct” reaction of thiol-ending monosaccharides with metal salts precursors. The formation of glyco-CuInS2 QDs could be explained by a nucleation-growth mechanism following the LaMer model. As-prepared four glyco-CuInS2 QDs were water-soluble, monodispersed, spherical in shape and exhibited size range of 3.0–4.0 nm. They exhibited well-separated dual emission in the visible region (500–590 nm) and near-infrared range (~ 827 nm), which may be attributable to visible excitonic emission and near-infrared surface defect emission. Meanwhile, the cell imaging displayed the reversibly distinct dual-color (green and red) fluorescence in tumor cells (HeLa, A549, MKN-45) and excellent membrane-targeting properties of glyco-CuInS2 QDs based on their good biorecognition ability. Importantly, these QDs succeed in penetrating uniformly into the interior (the necrotic zone) of 3D multicellular tumor spheroids (MCTS) due to their high negative charge (zeta potential values ranging from − 23.9 to − 30.1 mV), which overcame the problem of poor penetration depth of existing QDs in in vitro spheroid models. So, confocal analysis confirmed their excellent ability to penetrate and label tumors. Thus, the successful application in in vivo bioimaging of these glyco-QDs verified that this design strategy is an effective, low cost and simple procedure for developing green nanoparticles as cheap and promising fluorescent bioprobes.

Introduction

Nanobiosensor have played a key role in the fast cancer diagnosis and in clinical medicine and gained a great deal of attention in the areas of biological sciences [1, 2]. Recently, glyconanoparticles (GNPs) have been of increased interest as biofunctional nanosensors, which effectively combine the unique chemico-physical and optical properties of various nanoparticles with the characteristics of the carbohydrate coating [3,4,5]. Carbohydrates on the surface of NPs not only enhance the water solubility, biocompatibility and stability, but also endow these NPs with recognition ability. Thus, the GNPs have shown great potential for biomedical applications, especially in cellular labeling and imaging [6, 7]. Therefore, research on the biological effects associated with carbohydrate and developing new GNPs have become a hot topic in recent years.

A variety of nanoparticles (NPs), including gold, iron oxide, and semi-conductor quantum dots (QDs), have been used as the carriers of carbohydrates to prepare GNPs since the first synthesis of carbohydrate-functionalized gold nanoparticles in 2001[8]. Among them, QDs have been found as an attractive biomedical resource and widely used in vitro and in vivo for multiple color imaging and targeted drug delivery due to their broad absorption, high quantum yield, and long-term photostability [9,10,11]. Besides, many researches indicate that QDs have great potential in cancer detection and treatment [12, 13]. Good water solubility, biocompatibility, and biological targeting of QDs, all of which are essential for biomedical application, could easily be achieved through surface modification or conjugation with carbohydrates. Therefore, glyco-quantum dots (glyco-QDs) have attracted significant attention and many different glyco-QDs have been developed by coating with a variety of carbohydrates. For example, Shin-Ichiro Nishimura et al. synthesized a series of carbohydrate-capped Cd-based quantum dots (CdSe/ZnS and CdSeTe/CdS) with a variety of common sugars, such as α-glucose, α-mannose, lactose, N-acetyl-β-D-glucosamine, and N-acetyl-β-D-galactosamine. In vivo near-infrared (NIR) fluorescence imaging of these glyco-QDs revealed the importance of the terminal sialic acid residues for achieving prolonged in vivo lifetime [14]. Recently, Richichi's group reported a kind of CdSe/ZnS QDs-based fluorescent glyconanoprobe as nanoprobes for carbonic anhydrase IX imaging in cancer cells [15]. However, to date, the majority of glyco-QDs related research in biomedicine focused on Cd-based QDs including CdS, CdSe, and CdTe QDs, which are often added by ZnS and/or ZnSe as protective shells. Unfortunately, many studies have identified that highly cytotoxic Cd2+ ions in QDs can been released in cellular environment, which would ultimately increase the cytotoxicity effect of these QDs in a biological environment [16, 17]. So, the broad potential applications of Cd-based QDs in clinical medicine have been questioned by many biomedical researchers due to the high toxicity of heavy metal cadmium [18]. Accordingly, the potential high toxicity of fluorescent GNPs based on Cd-based QDs severely hinders their biomedical application. These results motivated the development of Cd-free glyco-QDs with less toxicity.

Currently, there has been a growing effort to prepare environmentally friendly Cd-free and less toxic QDs with performance comparable to or even better than existing Cd-based QDs [19]. Among these Cd-free QDs, copper indium sulfide (CuInS2) QDs have shown the greatest application potential for biological applications. CuInS2 QDs are I-III-VI2 semiconductor nanocrystals and does not contain any toxic heavy metals. They can provide PL emission ranging from the visible to the NIR, and have broad absorption, size-tunable photoluminescence (PL), excellent optical, and chemical stability [20, 21]. However, the conventional synthesis of CuInS2 QDs usually employs toxic organic solvents and the hydrophobic QDs are not appropriate for their application in biomedical fields. In addition, the use of organic solvents in the synthetic process is harmful to environment. In the view of hydrophilicity, toxicity and environment protection, the aqueous synthesis of high quality CuInS2 QDs is desired. In recent years, many efforts have been devoted to the aqueous synthesis of CuInS2 QDs for biological applications. Therefore, CuInS2 QDs could offer the opportunity to realize the potential of glyco-QDs without the toxicity limitations in biological environment. However, there is little research on the preparation and application of GNPs materials based on CuInS2 QDs in the biomedical field [22].

In general, the synthetic protocols for the preparation of glyco-QDs reported so far can be mainly described as: covalent conjugation, thiolated-glycans exchange, adsorption, and biotin-streptavidin chemistry [23]. Among them, valence conjugation is generally regarded as a straightforward and simple method. By following this route, functional carbohydrates are directly covalently linked on the surface of QDs. For instance, Penadés et al. directly used the thiol-ending saccharides as capping ligand and stabilizer to form the GNPs based on CdS QDs in degassed water under room temperature by a straightforward single step method [24]. Obviously, the direct synthesis of high-quality semiconductor glyco-QDs in aqueous medium is less harmful to the environment and health of people compared with the conventional organic synthesis of QDs. And, the resultant QDs are better water-soluble and biocompatible. Nevertheless, to the best of our knowledge, there is almost no literature on the synthesis of glyco-QDs based on ternary CuInS2 QDs in aqueous solution. Therefore, there is a high need to develop nontoxic and water-soluble glyco-CuInS2 QDs capped by carbohydrates as bio-compatible labeling and imaging probes in various biological applications.

Therefore, we developed in this work a simple, green and reproducible methodology for the construction of glyco-QDs in water via direct synthesis methods. These target glyco-QDs were synthesized by the “direct” reaction of thiol-ending monosaccharides (SH-fructose, SH-galactose, SH-mannose, SH-glucose) with metal salts precursors (CuCl2·2H2O, InCl3·4H2O and CS(NH2)2) in water. The as-prepared glyco-QDs not only exhibited well-separated dual-wavelength emission in the visible light region (500 ~ 590 nm) and the NIR range (~ 827 nm) upon excitation spectral scan (215 nm to 515 nm), but also displayed the high photostablity, good water-solubility, favorable photoreversibility and noncytotoxicity. Moreover, these glyco-CuInS2 QDs exhibited excellent biocompatibility and good biorecognition ability. The investigation of cellular imaging in tumor cells (HeLa, A549, MKN-45) indicated their reversible dual-color (green and red) imaging function and excellent membrane-targeting properties originating from a large number of carbohydrates on the surface of QDs. More importantly, they could penetrate uniformly into the interior of three-dimensional (3D) multicellular tumor spheroids (MCTS) due to their high negative charge, which realized the deep penetration depth of QDs in in vitro spheroid models. In summary, these glyco-CuInS2 QDs have successfully achieved 2D and 3D spatiotemporal visualization of the cell membrane, thus providing a huge potential in the application of commercial probes for cell membrane imaging.

Results and discussion

Design and synthesis of four kinds of glycol-CuInS2 QDs

The most common synthetic protocols for the preparation of glyco-QDs are ligand exchange with hydrophilic QDs or surface modification with carbohydrate [25,26,27]. However, the former requires the use of toxic organic solvents and the other one undergoes complex multistep process. Therefore, toxic organic solvent-free technology and simple method for hydrophilic synthesis of QDs are desired. Thus, we proposed a simple, green and reproducible design methodology of the construction of low cost glyco-QDs by the “direct” reaction of neutral monosaccharides (D-fructose and D-glucose mainly from honey, D-galactose from animal milk and D-mannose from fruit peels) with metal salts precursors in aqueous media. Scheme 1 illustrates the synthetic routes to four kinds of glycol-CuInS2 QDs, and the detailed synthesis procedure are described in Additional Information. The thiolation of monosaccharide is a key intermediate in the total synthesis of glyco-CuInS2 QDs. As shown in Scheme 1, four thiol-functionalized monosaccharides (SH-fructose, SH-galactose, SH-mannose, SH-glucose) were obtained by using DCC/DMAP esterification procedure with 3-mercaptopropionic acid. Subsequently, four environmentally friendly glyco-CuInS2 QDs were synthesized by using only these thiol-functionalized monosaccharides as stabilizing agents and capping ligands in aqueous medium by reaction with CuCl2·2H2O, InCl3·4H2O and CS(NH2)2 under an Ar atmosphere.

Scheme 1
scheme 1

Schematic illustration of the synthesis and biolabelling process of four kinds of glycol-CuInS2 QDs

In order to explore the optimal conditions for the synthesis of glyco-CuInS2 QDs, the molar ratios between Cu, In, S and the thiol-ending saccharide ligands, pH of the reaction solutions and reaction temperatures were systematically investigated. Additional file 1: Fig. S3–S5 show these changes in fluorescence intensity at 827 nm of the synthesized glyco-CuInS2 QDs at the different molar ratios, pH and reaction temperatures, respectively. Meanwhile, the optimal synthesis conditions are summarized in Additional file 1: Table S1. The experimental results show that the intensities of fluorescence of Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs reach the strongest when the ratios of Cu, In, S to ligands were 1:1:2:36, 1:1:2:24 or 1:1:2:12 at pH 11 ~ 12 and 150 °C, respectively. Moreover, the emission wavelength of these glyco-CuInS2 QDs do not change significantly with the change of the above reaction conditions. Thereupon, we synthesized a series of structurally well-defined glyco-CuInS2 QDs with natural carbohydrates on the surface.

Structural characterization of glyco-CuInS2 QDs

These synthesized glyco-CuInS2 QDs were characterized by TEM, DLS, XRD, FT-IR and XPS to demonstrate whether CuInS2 QDs were successfully encapsulated by the peripheral sulfhydryl monosaccharides, respectively. The effect of monosaccharide composition on crystal structure and morphology of CuInS2 QDs was first confirmed. As shown in the TEM images of glyco-CuInS2 QDs (Fig. 1 a1–a4), all monosaccharide-capped CuInS2 QDs have regular spherical shape and good monodispersity with mean diameters in the range of 3–4 nm in aqueous phase (Fig. 1b1–b4). Meanwhile, these insets to Fig. 1 a1–a4 show clear crystal lattice fringes with lattice spacing of 2.52 Å, 2.48 Å, 2.87 Å and 2.55 Å, respectively, which indicates the single crystal nature of the QDs. Figure 1c shows the XRD patterns of these synthesized glyco-CuInS2 QDs. The XRD spectra in Fig. 1c confirm the crystal structure of these glyco-CuInS2 QDs. The XRD pattern of glyco-CuInS2 QDs consisted of three major peaks with 2θ values of 27.9, 46.3, and 54.8, respectively, and all diffraction peaks corresponding to the (112), (024), and (116) indices of the tetragonal crystal structure (CuInS2, JCPDS no. 47-1372), respectively, suggesting that the crystal structure is well-maintained after capping with monosaccharides. According to the Debye–Scherrer formula: D = Kλ/Bcosθ (Where θ is the diffraction angle; λ is the X-ray wavelength; B is the half-height width of the diffraction peak; K = 0.89), the sizes of the prepared quantum dots were calculated to be 3.53 nm, 3.89 nm, 4.11 nm and 3.40 nm, respectively, which were consistent with the TEM and particle size test results. Furthermore, the clear lattice fringe (inset of Fig. 1a) indicated that the crystal structure of these CuInS2 QDs are not affected after capping with four monosaccharides. In addition, the broad diffraction peaks suggest that they have small sizes [28]. Besides, no other diffraction peaks were observed, which emphasizes the absence of any impurities in the prepared samples.

Fig. 1
figure 1

a1a4 TEM images of Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs at a scale bar of 5 nm, respectively (Inset: HR-TEM images of glyco-CuInS2 QDs and their lattice spacing). b1b4 Particle size distribution maps for Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs (Inset: Path diagram of glyco-CuInS2 QDs irradiated by laser pointer under daylight). c1 XRD diagram of Glu-CuInS2 QDs synthesized with four monosaccharides as ligands. d1d4 FTIR spectra of Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs for each of four sulfhydrylated monosaccharides and their synthetic glyco-CuInS2 QDs

Moreover, the IR spectrum reveal the composition of as-synthesized glyco-CuInS2 QDs. The FTIR spectra of the sulfhydrylated monosaccharides and sugar-capped CuInS2 QDs are compared in Figure d1-d4. The IR absorption bands of four sulfhydrylated monosaccharides at 3500–3000 cm−1 are ascribed to the stretching vibration of -OH, whereas the peaks at 3000–2900 cm−1, 2500–2600 cm−1 and 1750 cm−1 are assigned to -CH2-, -SH and C = O of the ester groups, respectively. Because the CuInS2 QDs are capped with monosaccharides, typical peaks of sugars can be found in the IR spectrum of glyco-CuInS2 QDs. More importantly, the almost disappearance of the -SH stretching vibrational peak indicates that these sulfhydrylated monosaccharides may combine on the surface of the CuInS2 through the -SH groups [29]. SEM–EDS spectrum of glyco-CuInS2 QDs had been performed in Additional file 1: Fig. S6 to further determine the elemental composition of these QDs. These results indicate that all of glyco-CuInS2 QDs contain C, O, Cu, In, and S elements. Meanwhile, the distribution of elements of CuInS2 QDs are also studied with EDX elemental mappings, the bright points indicate the high concentration of the elements in Additional file 1: Fig. S7. The results show that C, O, Cu, In, and S elements are homogeneously distributed throughout sample, suggesting that CuInS2 QDs are successfully capped with four monosaccharides, respectively.

Besides, XPS was used to further confirm the composition and the valence state of all elements structure in these monosaccharide-capped CuInS2 QDs (Additional file 1: Fig. S8). The XPS survey scans for glyco-CuInS2 QDs clearly displayed the C 1 s, O 1 s, Cu 2p, In 3d, and S 2p signals at around 285, 531, 952, 452 and 163.3 eV, respectively [30, 31]. To obtain deeper information about CuInS2 QDs and structure, the high-resolution XPS spectra were collected and analyzed. The high-resolution scan of the C 1 s spectrum (Fig. 2) can be fitted with four peaks at 284.1 (C-H), 284.6 (C–C), 285.5 (C–O) and 287.8 (C = O) eV [32, 33], and these results are in accordance with the results obtained by FTIR spectrum. Meanwhile, the high-resolution O 1 s spectra show the signals centered at 531 eV originating from the carboxyl oxygen (C = O) [34]. The Cu 2p core level is divided into two peaks representing Cu 2p3/2 (~ 932 eV) and Cu 2p1/2 (~ 951 eV), confirming the valence state of ions is + 1 rather than + 2 due to the disappearance of the peak at 944.0 eV for Cu2+ (Fig. 2) [35]. Therefore, the results indicated that the oxidation state of Cu2+ was reduced by sulfhydryl monosaccharide in the process of CuInS2 QDs preparation. Moreover, the valences states of In3+ (In 3d5/2, ~ 445 eV and In 3d3/2, ~ 452 eV) is also confirmed by XPS spectra [30]. Finally, the S is identified as -2 by analyzing the XPS spectra of S2p [35]. Thus, the existence of monosaccharide on the surface of CuInS2 QDs can be proved.

Fig. 2
figure 2

ad High-resolution XPS spectra of C 1 s, O 1 s, Cu 2p, In 3d and S 2p of Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs, respectively

Fig. 3
figure 3

Computational models a CuS2 model, b CuInS2 (001) model, c Fructose model, and d Fructose adsorbed on CuInS2 (001) model. Geometric optimization models e Fructose model, f Fructose adsorbed on CuInS2 (001) model, (g) CuInS2 (001) charge hair distribution; red: O, yellow: S, orange: Cu, brown: In, gray: C, black and gray: fixed atoms (bond length in Å; bond angle in)

Furthermore, it is important to establish the long-term structure and dimensional stabilities of QDs materials. Firstly, the zeta potentials of the synthesized CuInS2 QDs were determined by means of a zeta potential analyzer. According to previous reports [36], solutions with a zeta potential above + 20 mV and below − 20 mV were considered stable. As can be seen in Additional file 1: Fig. S9, the zeta potentials of the four ligand-synthesized QDs were − 30.1 mV, − 28.7 mV, − 23.9 mV, and − 25.2 mV, respectively, indicating that these glyco-CuInS2 QDs are stable in solution. It is noteworthy that there is a slight variation in the zeta potential values of these QDs when the surface charge density increases, indicating a reduced tendency to aggregate due to electrostatic repulsion [37]. In addition, the chemical properties of the four CuInS2 QDs were stable, as all the glyco-CuInS2 QDs still exhibited regular spherical morphology with an average diameter of about 4 nm after day fourteen, with good dispersion, thus ensuring the dimensional stability of these synthesized CuInS2 QDs (Additional file 1: Fig. S10).

Simulation study of Fru-CuInS2 QDs as an example

Calculation method

In this study, the DMol3 package in Material Studio was used to complete the calculations using a first-principles pseudopotential plane wave algorithm based on Koho-Sham self-consistent density generalization theory [38, 39]. The exchange–correlation energy was calculated using the generalized gradient approximation (GGA) and the Perdew-Burke-Ernzerhof (PBE) pseudopotential [40], according to which the interaction between the fructose adsorbed valence electrons and ions in the CuInS2 system can be described approximately. The cut-off energy is set at 489.8 eV for the optimization of the system structure, energy, density of states and energy band correlation calculations. The Brillouin zone integral uses a symmetric k-point method in Monkhors-Pack form, dividing the lattice according to 1 × 1 × 1. To avoid the interaction of repeated cycles of the surface in the z-axis direction, a vacuum layer of 10 Å is constructed above the surface. The convergence criteria for the electron self-consistent interaction energy, the force on all atoms and the maximum atomic displacement are 2 × 10–5 Ha, 0.004 Ha/Å and 0.005 Å respectively. These data were chosen as optimal values based on convergence verification.

Computational models

The tetragonal CuS2 with formation energy, density and band gap of -0.540 eV, 4.22 g/m3 and 0 eV were selected as 3.650 Å, 4.723 Å and 5.826 Å for a, b and c, as shown in Fig. 3a. The CuS2 doped in system with CuS2 (001) surface not only has a low surface energy but also is more easily exposed under typical adsorption reaction conditions. When Cu/In is 1/1, In provides the active site for fructose, which exhibits excellent catalytic performance as shown in Fig. 3b. Fructose is shown in Fig. 3c. In this study p (4 × 4) with 56 atoms were constructed using CuS2 as the substrate, with 12 In atoms doped on the surface of CuS2 (001) and cell sizes a, b and c of 11.599 Å, 11.599 Å and 15.799 Å, as shown in Fig. 3d.

Fig. 4
figure 4

a1d1 Dual fluorescence spectra of glyco-CuInS2 QDs at different excitation wavelengths; a2–d2 Dual fluorescence trend maps of glyco-CuInS2 QDs at different excitation wavelengths; a3d3 Fluorescence 3D images of glyco-CuInS2 QDs at different excitation wavelengths; a4d4 CIE chromaticity maps of glyco-CuInS2 QDs at different excitations

In the adsorption energy calculation, the atoms in the lowermost atomic layer are held in their initial positions in the equilibrium structure, while the top two atomic layers and the adsorbate are both released, and the system is geometrically optimized to reach a steady state [41], the adsorption energy (Eads) of Fructose adsorbed on the CuInS2 (001) surface is defined as:

$$E_{{{\text{ads}}}} = E_{{({\text{TG}})}} + E_{{({\text{Fru}})}} - E_{{({\text{Total}})}}$$

where: E(TG) is the total energy of the in doped CuS2 (001) system; E(Fru) is the energy of a single fructose molecule; E(Total) is the total energy of a fructose molecule adsorbed on the CuInS2 (001) system.

Calculation results and discussion

Firstly, the fructose molecular structure was optimized to obtain the most stable structure as shown in Fig. 3e. For the CuS2(001)-doped in system, the geometric optimization was performed to obtain the most stable optimization was performed to obtain the most stable condition of the system as shown in Fig. 3b. Then the adsorption of fructose molecules on the CuInS2(001) surface was considered, and the adsorption configuration was represented by Fru-CuInS2 as shown in Fig. 3f. The charge distribution map of the CuS2(001)-doped in system was calculated by optimization, as shown in Fig. 3g. As shown in Fig. 3, the optimized fructose molecules adsorbed on the surface of CuInS2(001) were obtained through simulations, and a significant change in the S–H bond length was found, with the bond length changing from 1.520 Å to 1.918 Å. This indicates that the fructose molecules adsorbed on the surface of CuInS2(001) were activated, resulting in chemisorption. Furthermore, the results of calculation of the adsorption energy of fructose adsorption on CuInS2 are shown in Table 1. The fructose adsorption in the CuInS2(001) system shows excellent adsorption performance with an energy of 3.05 Ha. The corresponding DOS and PDOS calculations for this system are shown in Additional file 1: Fig. S11. A comparison of the DOS plots of the two systems shows a sharp peak at -4.2 eV and a small peak at the Fermi energy level, indicating that the fructose molecules are adsorbed on the surface of CuInS2(001). A comparison of the PDOS plots of the two systems shows a new energy band at -23 eV-20 eV and a small peak at the Fermi energy level, indicating that the fructose molecules are adsorbed on the surface of CuInS2(001) with good adsorption properties.

Table 1 Calculation of the adsorption energy of fructose adsorption on CuInS2

Color-tunable emission of glycol-CuInS2 QDs

In order to understand the optical properties of these glyco-CuInS2 QDs and to open up a wider range of applications for these materials, their photophysical behavior has been investigated in depth. Firstly, UV–Vis absorption spectra of the as-prepared CuInS2 QDs are recorded in aqueous solutions (Additional file 1: Fig. S12). All of the glyco-CuInS2 QDs show strong absorption at 200–350 nm, assigned to the absorption bands of the thiolated monosaccharides. Besides, these UV–vis absorption spectra also show broad absorption from 400 to 800 nm without distinct excitonic absorption features, which is consistent with previous reports for QDs [42]. Therefore, the glyco-CuInS2 QDs exhibit wide absorption band covering the UV and Vis regions. Meanwhile, considerable absorption coefficient of the visible region is essential to exciting red or NIR fluorescence bioimaging in a confocal laser scanning microscopy (CLSM) system.

Subsequently, the photoluminescence (PL) behavior of four kinds of glyco-CuInS2 QDs were investigated at room temperature in detail. The emission spectra of glyco-CuInS2 QDs in aqueous solution with progressively longer excitation wavelengths from 215 to 535 nm in 20 nm increment are shown in Fig. 4a1–d1, respectively. It can be observed that all glyco-CuInS2 QDs exhibit two well-separated photoluminescence peaks in the visible light region (500 ~ 600 nm) and the NIR range (~ 827 nm) upon excitation spectral scan (215 nm to 515 nm), corresponding to the characteristic photoluminescence peaks of water-soluble CuInS2 QDs [43]. Recently, Arshad et al. also have reported an aqueous synthesis of CuInS2 QDs by using GSH as capping ligand, which showed two well-defined photoluminescence peaks at 550 nm and 725 nm [44]. The dual-emission phenomenon has a wide range of applications as chemosensors and markers for bioimaging [31, 45, 46]. Besides, glyco-CuInS2 QDs mainly emit NIR fluorescence at excitation wavelengths in the range 215–250 nm and green fluorescence in the range of 400–500 nm. As shown in Fig. 4a2-d2, these glyco-CuInS2 QDs show different green fluorescence emission maxima at 582 nm (Fru-CuInS2 QDs), 500 nm (Gal-CuInS2 QDs), 502 nm (Man-CuInS2 QDs) and 553 nm (Glu-CuInS2 QDs) under 515, 415, 395 and 455 nm excitation, respectively. And, all of glyco-CuInS2 QDs exhibit the same NIR fluorescence emission maxima at 827 nm under 215 nm excitation. The fluorescence quantum yields (QY) and lifetimes (τ) of these QDs for the green fluorescence emission in water were measured, respectively. As shown in Additional file 1: Table S2, the QYs of Fru-CuInS2, Gal-CuInS2 and Glu-CuInS2 QDs exceed 10% in water and Fru-CuInS2 QDs have highest QY (Φf = 15.05) compared with three other QDs (Φf,Gal-CuInS2 = 10.05%, Φf,Man-CuInS2 = 4.60%, Φf,Glu-CuInS2 = 12.10%). Furthermore, the average fluorescence lifetimes for the maximum emission of glyco-CuInS2 QDs in water are 4.27 ns, 4.71 ns, 3.96 ns and 6.61 ns, respectively (Additional file 1: Table S2 and Fig. S13). Unfortunately, the QYs and lifetimes of QDs for the NIR emission at 827 nm cannot be obtained due to the lack of suitable excitation source (~ 215 nm) in instrument testing.

Notably, the green and NIR emissions are well separated by > 235 nm, which is extremely beneficial for biosensing or bioimaging. Surprisingly, these glyco-CuInS2 QDs have very large Stokes shift of up to 612 nm for NIR emission (Additional file 1: Fig. S14), which allows for efficient separation of absorbance and emission maxima. More interestingly, the fluorescence emissions of glyco-CuInS2 QDs can be tuned from green to NIR by switching the excitation wavelength. As shown in Fig. 4a3-d3, lower-energy emission peak at 827 nm is much higher than the higher-energy emission peak around 500 nm under 215 nm excitation. And then, the green light intensity gradually increases and the NIR light intensity decreases with the increase in the excitation wavelength from 215 to 515 nm. Ultimately, the green fluorescence is in a dominant position. In addition, the approximate luminescence color changes of glyco-CuInS2 QDs were identified with the help of chromaticity color coordinates in the CIE chromaticity diagram (Fig. a4–d4). The CIE chromaticity diagram results indicate that different light colors4 are produced on the standard swatches at different excitation wavelengths in the range of 215 to 515 nm, while the light colors are excessive from red to green. Therefore, the luminescence colors of these glyco-CuInS2 QDs could be easily adjusted through changing the excitation wavelength.

Moreover, the photostability of glyco-CuInS2 QDs is extremely important feature used as fluorescence probes in biological staining and cell imaging. Therefore, the luminescent stabilities of four glyco-CuInS2 QDs have been investigated using fluorescence spectroscopy. As shown in Additional file 1: Fig. S15, all glyco-CuInS2 QDs in aqueous solution show good photostability because there are no obvious changes between the fluorescence spectra of glyco-CuInS2 QDs at the first day, the seventh days and the fourteenth days. Furthermore, light illumination stabilities of glyco-CuInS2 QDs in the aggregated state are also investigated under the illumination of two light sources and times. Firstly, all these solid-state glyco-CuInS2 QDs were prepared by adding dropwise the concentrated solutions on the glass slides for drying. Then, these glass slides were viewed under a fluorescence microscope. Significantly, all these solid-state QDs exhibited strong fluorescence signals irradiated by altering the excitation wavelengths in both of the green and red channels (Fig. 5). Besides, the luminescent brightness did not decrease significantly after 24 h of illumination. These test results also identify that all glyco-CuInS2 QDs have excellent light-stabilities.

Fig. 5
figure 5

Fluorescence microscopy images of glyco-CuInS2 QDs, obtained from green a and red b emission channels continuously irradiated for 0 h, 8 h, 16 h, 24 h respectively (the images also include the emission spectra of QDs at different excitation wavelengths)

Formation and dual emission mechanisms of glyco-CuInS2 QDs

The formation mechanism was summarized in Fig. 6. As shown in Fig. 6a, Cu+ and In3+ cations were released from metal precursors and coordinated with thiol-ending glycosides to form Cu-In thiolate [CuIn(SR)x]n after being mixed with thiol-ending monosaccharides (RS-H). Subsequently, [CuIn(SR)x]n was decomposed to QDs by thermolysis process. The formation of glyco-CuInS2 QDs could be explained by a nucleation-growth mechanism following the LaMer model [47]. Meanwhile, based on the hard-soft acid–base (HSAB) theory, Cl is hard base, while Cu+ and In3+ are soft and hard acid, respectively. So, Cu+ is much easier to be released from CuCl than In3+ ions from InCl3 due to the weaker interaction between Cu-Cl. On the other hand, RSH is soft base. So, the Cu-rich [CuIn (SR)x] n is formed because of the stronger coordinating ability between Cu+ and RS than that between In3+ and RS [48]. Thus, highly Cu-rich glyco-CuInS2 QDs are finally obtained with high density of surface defect states.

Fig. 6
figure 6

a Schematic illustration of formation mechanism of glyco-CuInS2 QDs; b Schematic diagram of semiconductor QDs photoluminescence (Solid lines represent radiative transitions, dashed lines represent non-radiative transitions)

Dual emission mechanism of glyco-CuInS2 QDs can further be elucidated based on their absorption and emission behavior. As shown in Fig. 6b, the energy band structure usually consists of a low-energy valence band (VB) filled with electrons and an empty high-energy conduction band (CB). The photon energy is transferred to the QDs if they can absorb incident light under laser excitation. Thus, under the action of high-energy rays, the electron jumps from the fundamental energy level of the valence band to the higher energy level of the conduction band. Subsequently, the photogenerated electrons are transferred nonradiatively to the intra-bandgap states (1). Meanwhile, the surface defects on the QDs capture some electrons, resulting in the intra-bandgap states (2). All electrons from two states undergo radiative and nonradiative recombination with holes [49]. Therefore, the emission monitored in the visible light region of 500 ~ 600 nm (high energy band) and the emission recorded at 827 nm (low energy band) may be attributable to visible excitonic emission and near-infrared surface defect emission, respectively. Indeed, this dual emission mechanism is consistent with previous published works on the two distinct transitions in CuxInS2 QDs [50, 51].

The pH-sensitive qualities of fluorescence

Due to the aggressive proliferation of tumor cells and the rapid formation of irregular blood vessels, nutrients and oxygen are rapidly consumed at the tumor site, causing lactic acid metabolites produced by glycolysis in the tumor cells to accumulate in the tumor interstitium, ultimately leading to a drop in the pH of the extra-tumor cellular environment to 6.2–6.9 [52]. Therefore, the high brightness of fluorescence probes in a weak acid solution is of great importance for in vivo imaging. Thus, it is essential to examine the effects of pH on the fluorescence intensity of glyco-CuInS2 QDs. As shown in Additional file 1: Fig. S16, the fluorescence of four glyco-CuInS2 QDs exhibit pH-sensitive response, and the maximum fluorescence intensity can be obtained when the pH is in the range of 6 to 7, which will be beneficial to the in vivo imaging.

The dual-color live-cell membrane imaging of glycol-CuInS2 QDs

Motivated by these excellent optical performances of glycol-CuInS2 QDs, we further study their imaging performances in living organisms. Three tumor cells (HeLa, A549, MKN-45) were selected for testing prior to performing cell imaging experiments. Firstly, the detection of cytotoxicity is a critical prerequisite for its application in cell imaging. Three kinds of tumor cells were incubated in concentrations of 20, 40, 60, 80 and 100 μg/mL of Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs aqueous solutions for 24 h, and their effects on cell viability were determined by MTT [53, 54], respectively. As shown in Additional file 1: Fig. S17, the survival rates of three tumor cells are above 90% for all culture concentrations, which indicated that all glycol-CuInS2 QDs possesses low cytotoxicity towards the living cells.

Besides, their abilities to enter tumor cells are also essential for biological imaging. Therefore, we investigated how glycol-CuInS2 QDs entered three kinds of tumor cells before performing cell staining. The dynamical interaction process of glycol-CuInS2 QDs with tumor cells at different time points through a live cell workstation (Olympus DP80) were recorded in Fig. 7a1–d1, respectively. It can be observed that the aggregated QDs particles slowly enter the cell membrane with time migration from 0 to 100 min. These results indicate that all glycol-CuInS2 QDs enter target cells by endocytosis, which occurs at the cell surface. The low cytotoxicities and excellent cell-entry abilities of these glycol-CuInS2 QDs provide strong in vivo evidences for live cell staining.

Fig. 7
figure 7

a1d1 the images of 0–100 min migration of aggregated glyco-CuInS2 QDs towards the cell membrane, respectively; a2a5 the imaging images of HeLa cells treated with 100 μg/mL glyco-CuInS2 QDs for 2 h, 6 h and 12 h, respectively; b2b5 the images of A549 cells treated with 100 μg/mL glyco-CuInS2 QDs for 2 h, 6 h and 12 h, respectively; c2c5 the imaging images of MKN-45 cells treated with 100 μg/mL glyco-CuInS2 QDs for 2 h, 6 h and 12 h, respectively (scale bar 10 μm); d2d5 mean brightness values of HeLa, A549, MKN-45 cells after incubation with glyco-CuInS2 QDs for 2 h, 6 h and 12 h

Subsequently, cell imaging experiments were performed to further demonstrate the practical application of four glycol-CuInS2 QDs. The HeLa, A549, MKN-45 cells were individually isolated in 12-well plates and incubated with glycol-CuInS2 QDs at a concentration of 100 μg/mL for 2, 6 and 12 h, respectively. As shown in Fig. 7, all glycol-CuInS2 QDs are highly concentrated in HeLa, A549, MKN-45 cell membrane and the fluorescence signals are continuously enhanced with the incubation time increased. For example, almost no fluorescent signals are observed in membrane of three tumor cells because only a few Fru-CuInS2 QDs enter cells after 2 h in Fig. 7a2–c2. With the entry of more QDs in cell membrane after 6 h incubation, the increasing fluorescent signals can be observed. Finally, significant fluorescent signals are observed after 12 h incubation, suggesting that Fru-CuInS2 QDs can be effectively captured by cell membrane. Meanwhile, Gal-CuInS2 QDs, Man-CuInS2 QDs and Glu-CuInS2 QDs also exhibit similar imaging abilities, respectively (Fig. 7a3–c3, Fig. 7a4–c4 and Fig. 7a5–c5). Furthermore, all glycol-CuInS2 QDs show same dual-colorful and tunable fluorescence signals by altering the excitation wavelengths in green and red emission channels.

Besides, methods of imaging provide the means to visualize in space and time, but there is still a great need to make quantitative measurements. Image J is an open-source image processing platform for multidimensional image data, built for the specific needs of scientific images [55]. In this paper, we made use of the Image J software toolkit, which has utility in scientific image analysis, to quantify the imaging fraction. For example, as shown in Fig. 7d2, the specific fluorescence intensities of HeLa, A549 and MKN-45 cells incubated with Fru-CuInS2 QDs are 1.450 a.u, 3.069 a.u. and 5.748 a.u after 2 h of incubation, and increase to 75.656 a.u., 59.634 a.u. and 64.705 a.u. after 6 h of incubation, and significantly increase to 243.413 a.u., 214.783 a.u. and 225.490 a.u. after 12 h of incubation, respectively. Meanwhile, the specific fluorescence intensities of HeLa, A549 and MKN-45 cells incubated respectively with Gal-CuInS2 QDs, Man-CuInS2 QDs and Glu-CuInS2 QDs also exhibit the similar growth trends with the increasing of incubation time (Fig. 7d3–d5).

Most notably, confocal imaging experiments show the excellent specific cytomembrane location for living tumor cells. According to a previously reported literature [56], based on the recognition of glyco-cell membrane receptor proteins, nanodots with more glycosyl modifications are more recognized by the receptor, entered the cells more and fluoresced brighter. Thus, these glycol-CuInS2 QDs can specifically identify cancer cells with glyco-receptor protein overexpression on the cell membrane for fluorescence imaging. The cell membrane is involved in various cellular processes and biological functions, such as cell signalling, cell adhesion, endocytosis, cytosol and selective permeation of substances, therefore, the cell membrane is indispensable to the cell, and its observation can also yield much information related to cellular state and disease. In view of this, the synthesis of cell membrane-targeted fluorescent bioprobes is of great importance [57]. Furthermore, in order to determine whether the fluorescence emitted by the two channels remains in the same position, a fluorescence co-localization assay was performed [58]. From the scattering trend of pixel points in the co-localized pixel plots in Fig. 8, we can find that the fluorescence of glycol-CuInS2 QDs emitted by the red and green channels are in a somewhat linear relationship and favor the red channel, indicating that the red and green fluorescence are co-localized in a certain proportion, but the green fluorescence is slightly darker than the red fluorescence. We then carried out co-localization analysis by a series of different parameters, Pearson's correlation coefficient (PCC) [59]. Figure 8a4–d4 shows the overlap rates of these glycol-CuInS2 QDs in different cell imaging with red and green channel imaging. Reassuringly, all overlaps are above 90%, indicating high targeting specificity.

Fig. 8
figure 8

a1d1 Pixel plots of imaging confocal after treatment of HeLa cells with Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs, respectively; a2d2 Pixel plots of imaging confocal after treatment of A549 cells with Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs, respectively; a3d3 Pixel plots of imaging confocal after treatment of MKN-45 cells with Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs, respectively; a4d4 Plots of threshold versus Pearson correlation coefficient in fluorescence confocal detection (inset: overlap of different channel imaging positions for the three cells)

The imaging of glycol-CuInS2 QDs in multicellular tumor spheroids

The studies of QDs imaging had been carried out in vitro with cancerous cell lines. However, only research in 2D monolayers model is unable to evaluate some very important abilities for glycol-CuInS2 QDs in cancer imaging, for example the ability of QDs to penetrate and retain within a tumor. At present, the challenge associated with QDs is their poor penetration depth, constraining them to a peripheral accumulation in in vivo tumors. Multicellular tumor spheroids (MCTS) have gained increased recognition as a useful three-dimensional (3D) tumor spheroid model for evaluating the effect of exogenous molecules on tumor progression in cancer research because they may closely mimic some physiological characteristics of solid tumors. However, to date, only few studies were conducted to investigate the QDs imaging in 3D in vitro spheroids models. Therefore, MCTS are used to evaluate the tumor penetration capacity of four glycol-CuInS2 QDs in this study, respectively (Additional file 1: Video S1–S4). Figure 9a shows the transport of glycol-CuInS2 QDs inside tumor spheroids at different time points, and a time-dependent penetration of glycol-CuInS2 QDs in multicellular tumor spheroids can been observed from 20 to 180 min. At 20 min post-treatment, red fluorescent spots in the confocal images indicate a distribution of glycol-CuInS2 QDs in the outer region of the 3D model. As the time increases, enhanced fluorescence signal in the inner region can be observed, suggesting more homogeneous uptake. Finally, there glycol-CuInS2 QDs penetrated almost to the center of the 3D spheroids after 180 min of incubation. Besides, as shown in Fig. 9b, the fluorescent signals of Gal-CuInS2 and Man-CuInS2 QDs are observed dispersed evenly throughout tumor cells, indicating the penetration capacities are higher than those of Fru-CuInS2 and Glu-CuInS2 QDs. Previous studies have shown that the internalization ability of QDs to HeLa cellular spheroids is affected by the QDs surface charges. In general, positively-charged QDs cannot reach the necrotic zone since they can be absorbed by the proliferating cells of the spheroid. While the negatively-charged QDs can diffuse more quickly, allowing them to penetrate deeply into the tissues [56, 60]. Thus, our results are in excellent agreement with the experimental results reported in the previous study. Figure 9c shows the schematic representation of glyco-CuInS2 QDs penetration in 3D MKN-45 spheroid. Glyco-CuInS2 QDs can succeed in penetrating uniformly into the interior of 3D MCTS and reach the necrotic zone due to their high negative charge (zeta potential values ranging from − 23.9 to − 30.1 mV). which overcame the problem of poor penetration depth of existing QDs in in vitro spheroid models. These results indicate that glycol-CuInS2 QDs have excellent tumor penetration capacity, which is of particular importance since the fluorescence probes used for cancer imaging must efficiently penetrate tumor tissues to reach all of the viable cells [61, 62]. Therefore, these glycol-CuInS2 QDs have great potential to be used as a fluorescent probe for intratumoral analysis.

Fig. 9
figure 9

a Fluorescence expression of glyco-CuInS2 QDs in 3D MKN-45 cell microspheres at different times (Scale bar: 500 µm); b the corresponding quantitative fluorescence intensity values; c Schematic representation of glyco-CuInS2 QDs penetration in 3D MKN-45 spheroid

The dual-color bioimaging of glycol-CuInS2 QDs in zebrafish

Finally, we investigated the imaging abilities of glycol-CuInS2 QDs in living organisms. Based on the genetic similarity of 87% to humans, the zebrafish has outstanding advantages as a model organism because its experimental.results are applicable to humans in most cases. Nowadays, zebrafish has become one of the most valuable models in vertebrate developmental biology and shows great potential for use in other disciplines [63]. Thus, using zebrafish, it is possible to study fundamental questions in the life sciences and to reveal the molecular mechanisms of embryonic and tissue organ development. Therefore, zebrafish is chosen as model organism in our study. Firstly, zebrafishes were incubated with glycol-CuInS2 QDs aqueous solution (1 μM) in a 96-well plate for 1 h. Subsequently, two-channel fluorescence signals were recorded by the fluorescence microscope. As shown in Fig. 10, all glycol-CuInS2 QDs in zebrafish display strong fluorescence signals in green and red channels under different excitation sources of fluorescence microscopy. The results indicate that these glycol-CuInS2 QDs can be absorbed through the skin of zebrafish and are low toxicity in biological systems. Moreover, we used Image J software to perform accurate calculations of the fluorescence intensity of the imaging (Additional file 1: Fig. S18) and used plug-ins to measure the fluorescence position in live imaging of different channels of zebrafish (Fig. 10e1-e4). The trends of pixel point dispersion in the QDs show the proportional overlap in the imaging of the two channels, and the overlap rate is quantified in Fig. 10f, which surprisingly remains above 91.56% for all glycol-CuInS2 QDs, showing a high degree of overlap. Therefore, these results clearly demonstrate that these glycol-CuInS2 QDs synthesized by aqueous method have great potential in in vivo bioluminescence imaging.

Fig. 10
figure 10

Microscopic images of two channels of zebrafish after 1 h incubation with glyco-CuInS2 QDs. a1d1 bright field; a2d2 green channel; a3d3 red channel; a4d4 merged. Scale bars are 600 μm. e1e4 Pixel plots of imaging confocal after treatment of zebrafish with Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs samples; (f) Plot of threshold versus Pearson's correlation coefficient in fluorescence confocal detection (inset: four samples treated with zebrafish overlap of the red and green channel imaging positions)

Conclusions

We developed an eco-friendly green approach to the aqueous-phase synthesis of QDs-based fluorescent nanobiosensor by the “direct” reaction of thiol-ending natural monosaccharides with metal salts precursors. Compared with existing synthesis methods through surface modification of QDs by carbohydrate, the “direct” strategy is simpler and more low-cost. The formation of glyco-CuInS2 QDs could be explained by a nucleation-growth mechanism following the LaMer model. As-prepared glyco-CuInS2 QDs (Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs) showed monodisperse spherical morphology (< 5 nm), good water solubility and non-cytotoxicity. These glyco-QDs exhibited excitation-wavelength-dependent visible/NIR dual emission with good photostablity and favorable photoreversibilit, which may be attributable to visible excitonic emission and near-infrared surface defect emission, respectively. Confocal microscopy experiments showed that glyco-QDs had highly specific targeting ability for cancer cell membranes (HeLa, A549, MKN-45) due to their good biorecognition ability originating from glycocluster on the surface of the CuInS2 QDs. Therefore, these glyco-QDs can be used as a potential plasma membrane imaging reagent for dual-color (green and red) imaging of cell membrane. More importantly, these glyco-QDs could succeed in penetrating uniformly into the interior of 3D multicellular tumor spheroids (MCTS) because of their high negative charge, thus suggesting the possibility to exploit such nanoprobes for the tumor bioimaging with excellent tumor penetration capacity. Therefore, our prepared glyco-QDs shows great commercial prospects in the field of biomedicine due to their low cost, simple manufacturing process and deep tumor penetration capacity. Besides, these glyco-CuInS2 QDs can exhibit more efficient recognition of tumor cells by coating with tumor-targeting agents such as cRGD.

References

  1. Nasrollahpour H, Khalilzadeh B, Hasanzadeh M, Rahbarghazi R, Estrela P, Naseri A, Tasoglu S, Sillanpää M. Nanotechnology-based electrochemical biosensors for monitoring breast cancer biomarkers. Med Res Rev. 2022;12:1–106.

    Google Scholar 

  2. Pourmadadi M, Rahmani E, Rajabzadeh-Khosroshahi M, Samadi A, Behzadmehr R, Rahdar A, Ferreira LFR. Properties and application of carbon quantum dots (CQDs) in biosensors for disease detection: a comprehensive review. J Drug Deliv Sci Tec. 2023;80: 104156.

    Article  CAS  Google Scholar 

  3. Sha Q, Fei J, Tu C, Liu B, Hu Z, Liu X. A universal strategy of glyconanoparticle preparation using a bifunctional linker for lectin sensing and cell imaging. Microchim Acta. 2022;189:1–11.

    Article  Google Scholar 

  4. Hao N, Neranon K, Ramström O, Yan M. Glyconanomaterials for biosensing applications. Biosens Bioelectron. 2016;76:113–30.

    Article  CAS  PubMed  Google Scholar 

  5. Norouzi F, Pourmadadi M, Yazdian F, Khoshmaram K, Mohammadnejad J, Sanati MH, Chogan F, Rahdar A, Baino F. PVA-based nanofibers containing chitosan modified with graphene oxide and carbon quantum dot-doped TiO2 enhance wound healing in a rat model. J Funct. 2022;13:300.

    CAS  Google Scholar 

  6. Tang Q, Huang G. Preparation and applications of glyconanoparticles. Int J Biol Macromol. 2018;116:927–30.

    Article  CAS  PubMed  Google Scholar 

  7. Zhang X, Huang G, Huang H. The glyconanoparticle as carrier for drug delivery. Drug Deliv. 2019;25:1840–5.

    Article  Google Scholar 

  8. Fuente JMDL, Barrientos AG, Rojas TC, Rojo J, Canada J, Frenandez A, Penades S. Gold glyconanoparticles as water-soluble polyvalent models to study carbohydrate interactions. Angew Chem Int Edit. 2001;40:2258–61.

    Google Scholar 

  9. Kim J, Hwang DW, Jung HS, Kim KW, Pham XH, Lee SH, Byun JW, Kim W, Kim HM, Hahm E, et al. High-quantum yield alloy-typed core/shell CdSeZnS/ZnS quantum dots for bio-applications. J Nanobiotechnol. 2022;20:22.

    Article  CAS  Google Scholar 

  10. Aires A, Möller M, Cortajarena AL. Protein design for the synthesis and stabilization of highly fluorescent quantum dots. Chem Mater. 2020;32:5729–38.

    Article  CAS  Google Scholar 

  11. Pourmadadi M, Moammeri A, Shamsabadipour A, Moghaddam YF, Rahdar A, Pandey S. Application of various optical and electrochemical nanobiosensors for detecting cancer antigen 125 (CA-125): a review. Biosensors. 2023;13:99.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Ashree J, Wang Q, Chao Y. Glyco-functionalised quantum dots and their progress in cancer diagnosis and treatment. Front Chem Sci Eng. 2020;14:365–77.

    Article  CAS  Google Scholar 

  13. Pourmadadi M, Rajabzadeh-Khosroshahi M, Saeidi Tabar F, Ajalli N, Samadi A, Yazdani M, Yazdian F, Rahdar A, Díez-Pascual AM. Two-dimensional grap hitic carbon nitride (g-C3N4) nanosheets and their derivatives for diagnosis and detection applications. J Funct. 2022;13:204.

    CAS  Google Scholar 

  14. Ohyanagi T, Nagahori N, Shimawaki K, Hinou H, Yamashita T, Sasaki A, Jin T, Iwanaga T, Kinjo M, Nishimura SI. Importance of sialic acid residues illuminated by live animal imaging using phosphorylcholine self-assembled monolayer-coated quantum dots. J Am Chem Soc. 2011;133:12507–17.

    Article  CAS  PubMed  Google Scholar 

  15. Biagiotti G, Angeli A, Giacomini A, Toniolo G, Landini L, Salerno G, Di Cesare ML, Ghelardini C, Mello T, Mussi S, Ravelli C, Marelli M, Cicchi S, Menna E, Ronca R, Supuran CT, Richichi B. Glyco-coated CdSe/ZnS quantum dots as nanoprobes for carbonic anhydrase IX imaging in cancer cells. ACS Appl Nano Mater. 2021;4:14153–60.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Allocca M, Mattera L, Bauduin A, Miedziak B, Moros M, De Trizio L, Tino A, Reiss P, Ambrosone A, Tortiglione C. An integrated multilevel analysis profiling biosafety and toxicity induced by indium- and cadmium-based quantum dots in vivo. Environ Sci Technol. 2019;53:3938–47.

    Article  CAS  PubMed  Google Scholar 

  17. Nikazar S, Sivasankarapillai VS, Rahdar A, Gasmi S, Anumol PS, Shanavas MS. Revisiting the cytotoxicity of quantum dots: an in-depth overview. Biophys Rev. 2020;12:703–18.

    Article  PubMed  PubMed Central  Google Scholar 

  18. Zikalala N, Parani S, Tsolekile N, Oluwafemi OS. Facile green synthesis of ZnInS quantum dots: temporal evolution of their optical properties and cell viability against normal and cancerous cells. J Mater Chem C. 2020;8:9329–36.

    Article  CAS  Google Scholar 

  19. Wang X, Tian J, Yong KT, Zhu X, Lin MCM, Jiang W, Li J, Huang Q, Lin G. Immunotoxicity assessment of CdSe/ZnS quantum dots in macrophages, lymphocytes and BALB/c mice. J Nanobiotechnol. 2016;14:10.

    Article  Google Scholar 

  20. Zhang Y, Qiao L, Zhang Z, Liu Y, Li L, Shen H, Zhao M. A mitochondrial-targetable fluorescent probe based on high-quality InP quantum dots for the imaging of living cells. Mater Des. 2022;219: 110736.

    Article  CAS  Google Scholar 

  21. Ding K, Jing L, Liu C, Hou Y, Gao M. Magnetically engineered Cd-Free quantum dots as dual-modality probes for fluorescence/magnetic resonance imaging of tumors. Biomaterials. 2014;35:1608–17.

    Article  CAS  PubMed  Google Scholar 

  22. Liu Z, Liu H, Wang L, Su X. A label-free fluorescence biosensor for highly sensitive detection of lectin based on carboxymethyl chitosan-quantum dots and gold nanoparticles. Anal Chim Acta. 2016;932:88–97.

    Article  CAS  PubMed  Google Scholar 

  23. Marradi M, Tricomi J, Matassini C, Richichi B. Carbohydrate functionalized quantum dots in sensing, imaging and therapy applications. In: Marco M, editor. Comprehensive Glycoscience. Amsterdam: Elsevier; 2021.

    Google Scholar 

  24. Jesús M, Penadés S. Glyco-quantum dots: a new luminescent system with multivalent carbohydrate display. Tetrah Asym. 2005;16:387–91.

    Article  Google Scholar 

  25. Krejcova L, Nejdl L, Rodrigo MM, Zurek M, Matousek M, Hynek D, Zitka O, Kopel P, Adam V, Kizek R. 3D printed chip for electrochemical detection of influenza virus labeled with CdS quantum dots. Biosens. 2013;54:421–7.

    Article  Google Scholar 

  26. Liu W, Greytak AB, Lee J, Wong CR, Park J, Marshall LF, Jiang W, Curtin PN, Ting AY, Nocera DG, et al. Compact biocompatible quantum dots via RAFT-mediated synthesis of imidazole-based random copolymer ligand. J Am Chem Soc. 2010;132:472–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Lima CN, Oliveira WF, Silva PMM, Filho PEC, Juul-Madsen K, Moura P, Vorup-Jensen T, Fontes A. Mannose-binding lectin conjugated to quantum dots as fluorescent nanotools for carbohydrate tracing. Methods Appl Fluoresc. 2022;10:025002.

    Article  Google Scholar 

  28. Branzi L, Purcell-Milton F, Cressoni C, Back M, Cattaruzza E, Speghini A, Gun’ko YK, Benedetti A. Chiral non-stoichiometric ternary silver indium sulfide quantum dots: investigation on the chirality transfer by cysteine. Nanoscale. 2022;14:12174–82.

    Article  CAS  PubMed  Google Scholar 

  29. Arshad A, Chen H, Bai X, Xu S, Wang L. One-pot aqueous synthesis of highly biocompatible near infrared CuInS2 quantum dots for target cell imaging. Chinese J Chem. 2016;34:576–82.

    Article  CAS  Google Scholar 

  30. Xiong W, Yang G, Wu X, Zhu J. Aqueous synthesis of color-tunable CuInS2/ZnS nanocrystals for the detection of human interleukin 6. ACS Appl Mater. 2013;5:8210–6.

    Article  CAS  Google Scholar 

  31. Guan X, Yang X, Lai S, Ding Y, Wei J, Zhang J, Zhang L, Li C, Tong J, Lei Z. Design and synthesis of biodegradable nonconjugated SSPAMAM dendrimers with unexpected deep-red/NIR emission and cell membrane targeting ability for biological imaging. Mater Des. 2022;221: 110982.

    Article  CAS  Google Scholar 

  32. Ding H, Zhang P, Wang T, Kong J, Xiong H. Nitrogen-doped carbon dots derived from polyvinyl pyrrolidone and their multicolor cell imaging. Nanotechnology. 2014;2: 205604.

    Article  Google Scholar 

  33. Viltres H, Odio OF, Biesinger MC, Montiel G, Borja R, Reguera E. Preparation of amine- and disulfide-containing PAMAM-based dendrons for the functionalization of hydroxylated surfaces: XPS as structural sensor. ChemistrySelect. 2020;5:4875–84.

    Article  CAS  Google Scholar 

  34. Zou W, Ma X, Zheng P. Preparation and functional study of cellulose/carbon quantum dot composites. Cellulose. 2019;27:2099–113.

    Article  Google Scholar 

  35. Liu J, Zhao X, Xu H, Wang Z, Dai Z. Amino acid-capped water-soluble near-infrared region CuInS2/ZnS quantum dots for selective cadmium ion determination and multicolor cell imaging. Anal Chem. 2019;91:8987–93.

    Article  CAS  PubMed  Google Scholar 

  36. Liu Z, Liu H, Liu L, Su X. Aptamer based lysozyme assay using fluorescent CuInS2 quantum dots and graphene oxide, and its application to inhibitor screening. Microchim Acta. 2016;183:2907–16.

    Article  CAS  Google Scholar 

  37. Prathna TC, Chandrasekaran N, Mukherjee A. Studies on aggregation behaviour of silver nanoparticles in aqueous matrices: effect of surface functionalization and matrix composition. Colloids Surf. 2011;390:216–24.

    Article  CAS  Google Scholar 

  38. Delley B. An all-electron numerical method for solving the local density functional for polyatomic molecules. J Chem Phys. 1990;92:508–17.

    Article  CAS  Google Scholar 

  39. Hammer B, Hansen LB, Nørskov JK. Improved adsorption energetics within density-functional theory using revised perdew-burke-ernzerhof functionals. Phys Rev B. 1999;59:7413–21.

    Article  Google Scholar 

  40. Perdew JP, Burke K, Ernzerhof M. Generalized gradient approximation made simple. Phys Rev Lett. 1996;77:3865.

    Article  CAS  PubMed  Google Scholar 

  41. Sun M, Peng Y. Study on structural, electronic and magnetic properties of Sn atom adsorbed on defective graphene by first-principle calculations. Appl Surf Sci. 2014;307:158–64.

    Article  CAS  Google Scholar 

  42. Yao J, Li P, Li L, Yang M. Biochemistry and biomedicine of quantum dots: from biodetection to bioimaging, drug discovery, diagnostics, and therapy. Acta Biomater. 2018;74:36–55.

    Article  CAS  PubMed  Google Scholar 

  43. Jiang T, Song J, Wang H, Ye X, Wang H, Zhang W, Yang M, Xia R, Zhu L, Xu X. Aqueous synthesis of color tunable Cu doped Zn-In-S/ZnS nanoparticles in the whole visible region for cellular imaging. J Mater Chem B. 2015;3:2402–10.

    Article  CAS  PubMed  Google Scholar 

  44. Arshad A, Akram R, Iqbal S, Batool F, Iqbal B, Khalid B, Khan AU. Aqueous synthesis of tunable fluorescent, semiconductor CuInS2 quantum dots for bioimaging. Arab J Chem. 2019;12:4840–7.

    Article  CAS  Google Scholar 

  45. Guan X, Lu B, Jin Q, Li Z, Wang L, Wang K, Lai S, Lei Z. AIE-active fluorescent nonconjugated polymer dots for dual-alternating-color live cell imaging. Ind Eng Chem Res. 2018;57:14889–98.

    Article  CAS  Google Scholar 

  46. Guan X, Li Z, Wang L, Liu M, Lei Z. Preparation of AIE polymer dots (Pdots) based on poly (N-vinyl-2-pyrrolidone)-Eu (III) complex and dual-color live cell imaging. Acta Chim Sinica. 2019;77:1268–78.

    Article  CAS  Google Scholar 

  47. Sowers KL, Swartz B, Krauss TD. Chemical mechanisms of semiconductor nanocrystal synthesis. Chem Mater. 2013;25:1351–62.

    Article  CAS  Google Scholar 

  48. Wang X, Liang Z, Xu X, Wang N, Fang J, Wang J, Xu G. A high efficient photoluminescence Zn–Cu–In–S/ZnS quantum dots with long lifetime. J Alloys Compd. 2015;640:134–40.

    Article  CAS  Google Scholar 

  49. Zaiats G, Kinge S, Kamat PV. Origin of dual photoluminescence states in ZnS–CuInS2 alloy nanostructures. J Phys Chem C. 2016;120:10641–6.

    Article  CAS  Google Scholar 

  50. Jara DH, Yoon SJ, Stamplecoskie KG, Kamat PV. Size-dependent photovoltaic performance of CuInS2 quantum dot-sensitized solar cells. Chem Mater. 2014;26:7221–8.

    Article  CAS  Google Scholar 

  51. Jara DH, Stamplecoskie KG, Kamat PV. Two distinct transitions in CuxInS2 quantum dots. bandgap versus sub-bandgap excitations in copper-deficient structures. J Phys Chem Lett. 2016;7:1452–9.

    Article  CAS  PubMed  Google Scholar 

  52. Jing X, Hu H, Sun Y, Yu B, Cong H, Shen Y. The intracellular and extracellular microenvironment of tumor site: the trigger of stimuli-responsive drug delivery systems. Small Methods. 2022;6:2101437.

    Article  CAS  Google Scholar 

  53. Zhang S, Zhang D, Ding Y, Hua J, Tang B, Ji X, Zhang Q, Wei Y, Qin K, Li B. Bacteria-derived fluorescent carbon dots for highly selective detection of p-Nitrophenol and bioimaging. Analyst. 2019;144:5497–503.

    Article  CAS  PubMed  Google Scholar 

  54. Maddalena S, Walter P, Anna B, Antonella P, Markus D, Lorenza M, Rita C. Manganese in diagnostics: a preformulatory study. Pharmaceutics. 2022;14:108.

    Article  Google Scholar 

  55. Arena ET, Rueden CT, Hiner MC, Wang S, Yuan M, Eliceiri KW. Quantitating the cell: turning images into numbers with ImageJ. Wires Dev Biol. 2017;6: e260.

    Article  Google Scholar 

  56. Liu Y, Ji D, Dong L, Galanos N, Zang Y, Li J, Vidal S, He X. Supramolecular assembly of fluorogenic glyco-dots from perylenediimide-based glycoclusters for targeted imaging of cancer cells. ChemComm. 2017;53:11937–40.

    CAS  Google Scholar 

  57. Wang D, Su H, Kwok RT, Hu X, Zou H, Luo Q, Lee MMS, Xu W, Lam JWY, Tang BZ. Rational design of a water-soluble NIR AIEgen, and its application in ultrafast wash-free cellular imaging and photodynamic cancer cell ablation. Chem Sci. 2018;9:3685–93.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Naya G, Joseph BW, Paul RM, Edward JS, Johannes PM, Iestyn P, Lukas P, Cameron A, Arwyn TJ, Wolfgang L, Peter W, Paola B. Four-wave-mixing microscopy reveals non-colocalisation between gold nanoparticles and fluorophore conjugates inside cells. Nanoscale. 2020;12:4622–35.

    Article  Google Scholar 

  59. Xu W, Lee M, Nie J, Zhang Z, Kwok R, Lam J, Xu F, Wang D, Tang B. Three-pronged attack by homologous far-red/NIR AIEgens to achieve 1+1+1>3 synergistic enhanced photodynamic therapy. Angew Chem Int Ed. 2020;132:9697–703.

    Article  Google Scholar 

  60. Dirheimer L, Pons T, Marchal F, Bezdetnaya L. Quantum dots mediated imaging and phototherapy in cancer spheroid models: state of the art and perspectives. Pharmaceutics. 2022;14:2136.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Przysiecka J, Michalska M, Nowaczyk G, Pepliſska B, Jesionowski T, Schneider R, Jurga S. RGD peptide as effective transporter of CuInZnxS2+x quantum dots into human cancer cells. Colloid Surface. 2016;146:9–18.

    Article  CAS  Google Scholar 

  62. Jiang X, Xin H, Gu J, Xu X, Xia W, Chen S, Xie Y, Chen L, Chen Y, Sha X, Fang X. Solid tumor penetration by integrin-mediated pegylated poly (trimethylene carbonate) nanoparticles loaded with paclitaxel. Biomaterials. 2013;34:1739–46.

    Article  CAS  PubMed  Google Scholar 

  63. Ko SK, Chen X, Yoon J, Shin I. Zebrafish A good vertebrate model for molecular imaging using fluorescent probes. Chem Soc Rev. 2011;40:2120–30.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Natural Science Foundation of China (Nos. 21965032; 22267012; 21761032; 52162034), the Natural Science Foundation of Gansu Province, China (Nos. 20JR5RA525; 20JR10RA143) and Doctoral Program Fund of Lanzhou University of Arts and Sciences (Nos. 2020BSZX08).

Author information

Authors and Affiliations

Authors

Contributions

XG: Conceptualization, writing – review & editing. LZ: Investigation, Writing – original draft. SL: Methodology, Funding acquisition. JZ: Data curation. JW: Validation. KW: Data curation. WZ: Investigation. CL: Visualization. JT: Project administration. ZL: Supervision. All authors read and approved the final manuscript.

Corresponding authors

Correspondence to Xiaolin Guan or Shoujun Lai.

Ethics declarations

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Supplementary Information

Additional file 1: Figure S1

1H NMR hydrogen spectra of (a) Fru-CuInS2 QDs, (b) Gal-CuInS2 QDs, (c) Man-CuInS2 QDs and (d) Glu-CuInS2 QDs (room temperature, D2O). Figure S2. MS spectra of (a) Fru-CuInS2 QDs, (b) Gal-CuInS2 QDs, (c) Man-CuInS2 QDs and (d) Glu-CuInS2 QDs. Figure S3. Fluorescence emission spectra of (a) Fru-CuInS2 QDs, (b) Gal-CuInS2 QDs, (c) Man-CuInS2 QDs, and (d) Glu-CuInS2 QDs synthesized at different dose ratios. (Inset: relationship between dose ratio and luminescence intensity of CuInS2 QDs). Figure S4. Fluorescence emission spectra of (a) Fru-CuInS2 QDs, (b) Gal-CuInS2 QDs, (c) Man-CuInS2 QDs, and (d) Glu-CuInS2 QDs synthesized under different pH conditions. (Inset: relationship between pH and luminescence intensity of CuInS2 QDs). Figure S5. Fluorescence emission spectra of (a) Fru-CuInS2 QDs, (b) Gal-CuInS2 QDs, (c) Man-CuInS2 QDs, and (d) Glu-CuInS2 QDs synthesized under different temperature conditions. (Inset: Relationship between temperature and luminescence intensity of CuInS2 QDs). Figure S6. The SEM–EDS analysis of Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs. Figure S7. The EDX elemental mapping of Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs. Figure S8. The XPS survey spectra of (a)Fru-CuInS2 QDs, (b)Gal-CuInS2 QDs, (c)Man-CuInS2 QDs, and (d)Glu-CuInS2 QDs. Figure S9. Zeta potential values of glyco-CuInS2 QDs. Figure S10. TEM images of Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs at a scale of 5 nm after fourteen days. Figure S11. DOS and PDOS of Fructose molecules adsorbed on CuInS2(001) before and after (a) DOS of CuInS2(001) system, (b) Fructose molecules adsorbed on DOS of CuInS2(001) system, (c) PDOS of CuInS2(001) system, (d) Fructose molecules adsorbed on DOS of CuInS2(001) system. Figure S12. The UV–Vis absorption spectra of Fru-CuInS2, Gal-CuInS2, Man-CuInS2 and Glu-CuInS2 QDs, respectively. Figure S13. Fluorescence lifetimes of glyco-CuInS2 QDs. Figure S14. Excitation and emission fluorescence spectra of (a) Fru-CuInS2 QDs, (b) Gal-CuInS2 QDs, (c) Man-CuInS2 QDs, and (d) Glu-CuInS2 QDs. Figure S15. Fluorescence stabilities of (a) Fru-CuInS2 QDs, (b) Gal-CuInS2 QDs, (c) Man-CuInS2 QDs and (d) Glu-CuInS2 QDs in solution. The black, red and blue lines refer to the fluorescence spectra of glyco-CuInS2 QDs after 0, 7, and 14 days, respectively. Figure S16. Fluorescence spectra of the pH response of (a) Fru-CuInS2 QDs, (b) Gal-CuInS2 QDs, (c) Man-CuInS2 QDs, and (d) Glu-CuInS2 QDs. (Inset: relationship between pH and luminescence intensity of CuInS2 QDs). Figure S17. MTT assay to detect changes in cellular activity of cancer cells after the addition of 20,40,60,80,100 μg/mL of (a) Fru-CuInS2 QDs, (b) Gal-CuInS2 QDs, (c) Man-CuInS2 QDs, and (d) Glu-CuInS2 QDs, respectively. Figure S18. Fluorescence intensities of (a)Fru-CuInS2 QDs, (b) Gal-CuInS2 QDs, (c) Man-CuInS2 QDs, and(d) Glu-CuInS2 QDs imaged under fluorescence microscopy after incubation of zebrafish for 1 h, respectively. Table S1. Exploration of optimal synthesis conditions for glyco-CuInS2 QDs. Table S2. The photophysical properties of glyco-CuInS2 QDs.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Guan, X., Zhang, L., Lai, S. et al. Green synthesis of glyco-CuInS2 QDs with visible/NIR dual emission for 3D multicellular tumor spheroid and in vivo imaging. J Nanobiotechnol 21, 118 (2023). https://doi.org/10.1186/s12951-023-01859-6

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12951-023-01859-6

Keywords