Pushpendra Singha,
Rajan Kumar Singh*ab and
Ranveer Kumar*a
aDepartment of Physics, Dr Harisingh Gour Central University, Sagar, 470003, M. P., India. E-mail: ranveerssi@yahoo.com; rajanphysicssgo@gmail.com; Tel: +91 9425635731
bDepartment of Chemical Engineering, National Taiwan University, Taipei, Taiwan, ROC
First published on 12th January 2021
Currently, developments in the field of quantum dots (QDs) have attracted researchers worldwide. A large variety of QDs have been discovered in the few years, which have excellent optoelectronic, antibacterial, magnetic, and other properties. However, ZnO is the single known material that can exist in the quantum state and can hold all the above properties. There is a lot of work going on in this field and we will be shorthanded if we do not accommodate this treasure at one place. This manuscript will prove to be a milestone in this noble cause. Having a tremendous potential, there is a developing enthusiasm toward the application of ZnO QDs in diverse areas. Sol–gel method being the simplest is the widely-favored synthetic method. Synthesis via this method is largely affected by a number of factors such as the reaction temperature, duration of the reaction, type of solvent, pH of the solution, and the precipitating agent. Doping enhances the optical, magnetic, anti-bacterial, anti-microbial, and other properties of ZnO QDs. However, doping elements reside mostly on the surface of the QDs. The presence of doping elements inside the core is still a major challenge for doping techniques. In this review article, we have focused on pure, rare-earth, and transition metal-doped ZnO QD properties, and the various synthetic processes and applications. Quantum confinement effect is present in nearly every aspect of the QDs. The effect of quantum confinement has also been summarized in this manuscript. Furthermore, the doping of rare earth elements and transition metal, synthetic methods for different organic molecule-capped ZnO QDs, mechanisms for reactive oxygen species (ROS) generation, drug delivery system for cancer treatment, and many more application are discussed in this paper.
Theoretically, QDs are crystals that have physical dimensions comparable to the exciton Bohr radius of the material of which they are made. These particles are confined in all the three dimensions. In case of ZnO QDs, it is found that the exciton Bohr radius is very small, which is nearly 0.9 nm. Thus, it is very difficult to synthesize ZnO QDs having radius less than the Bohr radius. Most of the ZnO QDs have a radius greater than the Bohr radius but comparable to it. These particles can be considered as quantum dots as they successfully show the effect of quantum confinement. It has an optical band gap of 3.37 eV, which further increases with the decrease in the particle size and large exciton binding energy (60 meV).8
Most of the semiconductor quantum dots are in the form of colloidal solution. Among other colloidal systems, the colloidal solution of ZnO QDs is in great demand as they have a large area of application, excellent safety, good biocompatibility, non-toxicity, and low cost. Also, its anti-bacterial activity, biocompatibility, reliable mechanical properties, and physicochemical stability makes it a desirable candidate for dental materials. It is a potential candidate as a disinfectant and an antibacterial agent. These properties depend upon the structural morphologies and defects, surface functionalization, and exposure conditions. An illustrative diagram (Fig. 1) is provided for detailed information about the synthesis, properties, and applications of ZnO QDs. Hence, ZnO shows a wide range of applications in material engineering, health science,9 drug delivery10,11 photocatalyst, gas sensors,12 cancer therapy13–15 photoelectric light-emitting diodes (LED),16–18 solar cells,19 and photo-detectors.20
Fig. 1 An overview of the synthesis of pure, doped, and conjugated ZnO, change in the properties, and their applications. |
Quantum confinement21–23 is a major aspect of QD that keeps it at the next level compared to nanoparticles. It has a direct influence over the optical properties of ZnO QDs.24 Three-dimensional quantum confinement of the charge carriers increases the life-time of the carriers and the photoluminescence intensity, which finally enhances the optoelectronic properties of the QDs. The quantum confinement effect can be easily understood as a “particle-in-a-box”. This phenomenon is observed when the electron wave-function is influenced by the size of the particles. For ZnO QDs having size less than 3.6 nm, a strong confinement is observed.25 To study the confinement effect practically, we need a 3D model. For such cases, “particle-in-a-box” is replaced by the “particle-in-a-sphere” model. In such cases, the expression for exciton binding energy can be written as
(1) |
Fig. 2 TEM, HRTEM, and SAED pattern of the ZnO QDs synthesized at (a) room temperature, (b) 40 °C, (c) 60 °C, and (d) 80 °C. (e and f) Photoluminescence and absorbance spectra of the ZnO QDs at different synthesis temperatures.29 (g) Training (blue) and testing (green, yellow and red) dataset of ZnO QDs. Adapted with permission from ref. 29. Copyright (2020) Elsevier. |
Quantum dots give excellent fluorescence as compared to nanoparticles due to the availability of the electron–hole pair to interact with the surface states of the quantum dots. The ZnO QDs significantly increase the fluorescence of the material as compared to the ZnO nanoparticles (Fig. 3e–g). ZnO QDs give yellow emission in the presence of UV light. The emission spectra may shift depending upon the precipitating agent used in the process, solvent,30 induced defects, and the size of the particles.31 The optical band gap of the QDs can also be tailored by the doping of different elements32,33 and by using capping agents. From Fig. 3a and b, we can see that by capping QDs with SiO2, the UV spectra gives a blue shift in the range from 340 nm to 310 nm, whereas on capping nanoparticles with SiO2, there is a red shift. This is attributed to the change in the particle size, which decreases upon capping in the QDs and increases in the nanoparticles (Fig. 3c and d). This is due to the restriction on agglomeration of ZnO QDs; the SiO2 coating plays the role of restricting the agglomeration. Liu et al.34 used Gd to reduce the size and increased the specific surface area of the ZnO QDs. Sun et al.35 found that besides the reduction in size and increase in the vacancy defects, La3+ also weakens the Zn–O bond, which give rise to more defects. However, excessive concentration of doping elements reduces the fluorescence emission intensity and the quantum yield.36 QDs also show the quenching of fluorescence intensity with the addition of the metal ion.37 The quenching of the fluorescence intensity can also be possible by the passivation of surface defects by the organic ligands. It was found that the intensity of the photoluminescence (PL) spectra of the doped QDs first increases with the concentration of the doping element and then starts decreasing with a further increase in the concentration of the doping elements. This is due to the weakening of the Zn–O bond with the increase in the doping concentration at the initial stage. This results in the production of more and more vacancy defects, which are responsible for the PL emission spectra. With a further increase in the concentration of the doping element, after a certain limit, the Zn–O bond breaks and the PL intensity starts quenching.38,39 Thus, ZnO QDs can be used as the fluorescence probe for the detection of metal ion impurities in drinking water as the detection limit for several ions is very low.40
Fig. 3 (a) and (b) UV-visible absorption spectra of SiO2-capped ZnO QDs and nanoparticles, respectively, (c and d) TEM images of SiO2-capped ZnO QDs and nanoparticles, (e) schematic illustration of the increase in the fluorescence spectra of a material by the QDs and the nanoparticles, (f and g) and variation in the fluorescence intensity.41 Adapted with permission from ref. 41. Copyright (2020) MDPI. |
ZnO QDs show strong anti-microbial properties at a specific pH42 as ZnO can be easily dissolved into Zn2+. Protected (QDs with any capping agent) QDs release Zn2+ ion at a specific location depending upon the capping agent.43 Thus, these are helpful in targeted drug delivery systems44 for the treatment of different diseases. Moreover, the concentration of the conduction band electron (e−) and valence band holes (h+) is very high even in the absence of UV (ultra-violet) light.45 The presence of more electrons and holes in the conduction and valence band, respectively, as compared to the QDs nanoparticles, shows more ROS (reactive oxygen species) generation. Hence, QDs have a large capacity to degrade organic molecules, microbes, and bacteria.46
The effect of the particle size on the Raman spectra of ZnO was also found.47 In bulk ZnO, there are two phonon modes: longitudinal optical (LO) and transverse optical (TO).48 These modes further split into A1 and E1 symmetries. There are two non-polar Raman-active phonon modes with E2 symmetries also present. The low frequency E2 mode is related to the vibration of the zinc ion lattice and the high frequency E2 mode is related to the vibration of the oxygen ion lattice. The phonon peak shift that arises in the ZnO QDs is related to three main factors: phonon localization by defect creation, confinement effect within the QD boundaries, and laser-induced heating effect.49–52 Alim and his group53 studied the resonant and non-resonant Raman spectra of ZnO. They found that the E2 (high) peak shifted by 3 cm−1, which was due to the presence of defects. These defects are found to increase in the QDs and doped QDs in a large amount. They had also found that the resonant peak of the LO phonon mode shifts linearly towards the lower frequency side with the increase in the UV laser power. From Fig. 4a and b, we can see that there is a small shift of 4 cm−1 in the LO phonon mode for bulk ZnO to ZnO QDs. A change in the laser power shows a large red shift (in tens of cm−1) as compared to the other two factors (in few cm−1).
Fig. 4 Resonant Raman spectra of (a) bulk ZnO nanoparticles and (b) ZnO QDs.53 Adapted with permission from ref. 53. Copyright (2005) AIP. |
Since the past decade, there has been a surge in the field of ZnO QDs. This may be the outcome of congestion in the branch of nanoparticles, where researchers have not left any stone unturned. Yet we have not found any sizeable review article that can help to steer the research work forward. To overcome this drawback, we have tried to assemble all the material related to ZnO QDs in this review. In this review we have mainly focused on ZnO QDs. We have tried to cover most of the work done by researchers worldwide in this field in the last few decades. The main parts of this review are the synthetic approaches, doped ZnO QDs, and applications. We have added the effect of quantum size on the different physical properties of ZnO in the introduction part itself. After introduction, we discuss the synthetic approaches adopted by different researchers for the fabrication of ZnO QDs. This part covers the various ZnO QD fabrication methods, which includes the method for water and ethanol stable QDs. We have also tried to cover different synthetic methods, which include the capping of ZnO QDs so that the QDs remain stable for a large duration. We have also studied the synthetic methods for the doping of different elements. In the next section, we have studied the effect of doping in the QDs. Here, our main focus was on rare earth-doped QDs, transition metal-doped QDs, and composite QDs. We have also covered the effect of doping on various physical properties. The last section is the application part. Here, we have studied the fabrication of some devices in brief. We have also focused on some other application besides the devices. The main content of this section is bio-sensors, photodetectors, light emitting diodes (LED), catalytic application, as well as anti-cancer and anti-bacterial applications.
The reaction temperature and time also affect the particle size. Chen et al.61 studied the effect of reaction temperature and time. They used water–ethanol mixed solvent in the 2:1 ratio for the preparation of the ZnAc solution. Then, they mixed NaOH solution at 50, 60, 70, 80, and 90 °C for 6, 7, 8, 9, and 10 h reaction time. From both these variations, they have found that the particle size increases with the reaction temperature and time. The PL (photoluminescence) intensity of the QDs after 7 h of reaction was found to be the maximum. According to Chen,61 it was the time required to complete the reaction. Similarly, they found a higher emission peak for 60 °C reaction temperature. Regonia29 also studied the effect of reaction temperature and time on the optical properties of ZnO QDs. They found an increase in the particle size and decrease in the band gap. The increase in the particle size was confirmed by the TEM images of the QD grown at different temperatures. We can see that the particle size increases for synthesis at room temperatures, 40, 60, and 80 °C. The UV-visible absorption and photoluminescence spectra of the samples confirmed the decrease in the band gap energy. This variation can be easily defended by the Ostwald ripening theory as discussed earlier. They also studied the effect of different synthetic conditions on the physical properties of the QDs. Machine learning has proven to be the best supporting tool for predicting different physical properties. Instead of using artificial neural networks (ANN), they used Kernel ridge regression (KRR) and ridge regression (RR) as it needs very limited data sets for training the algorithm in comparison to ANN, which takes a large amount of data for training. In the field of nanomaterials, it is very tedious to take such a large amount of data. From Fig. 7 we can see that the KRR and RR model performs much better than the ANN model. The predicted band gap is very near to the experimental value.
A modified sol–gel method known as E. Meulenkamp's method can be adopted for the synthesis of ZnO QDs. This produces uniform and mono-disperse ZnO QDs by the precise control of the water content. In this method, first, a precursor solution of zinc salt (mostly zinc acetate) was formed in dehydrated ethanol and then to produce QDs from it, a calculated amount of water was added. Chen and his group62 studied the effect of the water content on the growth kinetics of ZnO QDs. The focus of these studies was on the two main processes, i.e., oriented attachment (OA) and Ostwald ripening (OR). These result show that the water content affects both the processes. The TEM images of the QDs formed by this method are shown in Fig. 6. Fig. 6a shows the mono-disperse and uniform particles, and the SAED pattern in the inset. The crystal fringes shown in the inset of Fig. 6b confirm the 002 plane of ZnO (Table 1).
Fig. 6 TEM images of ZnO QDs. (a) SAED pattern in the inset and (b) HRTEM in the inset.62 Adapted with permission from ref. 62. Copyright (2019) Elsevier. |
Fig. 7 ML models for the band gap of ZnO QD at different temperatures and times.29 Adapted with permission from ref. 29. Copyright (2020) Elsevier. |
S. no. | Synthetic method | Solvent | Dopant | Capping agent | Composite materials | Change in properties | Ref. |
---|---|---|---|---|---|---|---|
1 | Wet chemical | Ethanol | Co, Mn | — | — | Size 2.34 to 3.46 nm | 63 |
2 | Reflux at 80 °C for 2 h | Ethanol | La, Co | KH-560 | — | The QY at 495 nm increases from 30.5% to 77.9% | 35 and 64 |
3 | Sol–gel | Ethanol | — | PEG | — | Selectivity toward the Cu2+ ion | 65 |
4 | Facile low-temperature solution process | Ethanol | — | PVP | — | Decomposes methyl orange | 66 |
5 | Solution method | Methanol and dispersed in water | — | TEOS | — | 4–8 nm size, blue shift in PL emission | 67 |
6 | Ultrasonication microreactor | Ethanol | — | PEG-400 | — | 2–3 nm size, green emission, quantum yield 64.7% | 68 and 69 |
7 | One pot method | Ethanol | — | NIPAM, TBAM, APM | Labelling of E. coli, water stable for 15 days | 70 | |
8 | Solution-based method | Ethanol | — | MSA | — | ∼3 nm size, white light emitting | 71 |
9 | Wet chemical method | Ethanol | — | — | — | 3–11 nm size | 72 |
10 | Sol–gel | Ethanol | — | — | — | 6–10 nm size | 73 |
11 | Pulsed laser deposition (PLD) method and rapid thermal annealing | — | — | — | — | Average size 10 nm | 74 |
12 | Sonication | Ethanol | — | — | PMMA | Stable and flexible | 75 |
13 | Sol–gel | Ethanol | — | TEOS | — | White light emission | 76 |
14 | Modified sol–gel | Ethanol as the solvent and water for the QDs from the precursor solution | — | — | — | Water controls the QDs growth | 62 |
15 | Solution based method | Ethanol | — | KAS, EDC, APTS | — | pH triggered antimicrobial activity | 77 |
16 | Sol–gel | Ethanol | — | CTAB, TEOS | — | ∼5 nm size, stable in saline water | 78 |
17 | Sol–gel | Ethanol | — | SBS | — | Stable under saline water for more than 30 days | 79 |
18 | Refluxed method | Ethanol | — | — | — | Cytotoxicity towards the MCF-7 and MDA-MB-231 cells | 80 |
19 | Sol–gel | Ethanol | — | Oleic acid and TMAH | PETTA and 2EEEA | Photooxidative degradation at selective sites | 81 |
20 | Hydrothermal | Ethanol–water mixed solvent | — | — | CuO sheet | Photocatalytic and antibacterial activity | 82 |
21 | Sol–gel | Ethanol | Pr3+ | — | — | Energy transfers from the Pr3+ ions to the ZnO QDs. | 32 |
22 | Sol–gel | Ethanol | Mn2+ | TOPO | — | Ferromagnetism | 83 |
23 | Sol–gel | Ethanol | Rare earth element | — | — | Photocatalytic and photoluminescence | 84 |
24 | Sol–gel | Ethanol | Cu | — | — | Increases field emission by nanorods | 39 |
25 | Ultrasonic method | Ethanol | tin | — | — | Changes direct band gap to indirect band gap | 85 |
Geng et al.65 used polyethylene glycol (PEG) for the preparation of water-stable monodisperse PEG capped ZnO QDs (shown in Fig. 8f). First, they prepared ZnO QDs in ethanol using ethyl acetate as the precipitating agent. Then, they added 4 mL PEG in 4 mL ZnO QDs and churned it at room temperature for 60 min. Finally, after centrifugation and washing with ethanol and ultrapure water, the QDs were obtained by dispersing them in 10 mL ultrapure water. Similarly, Rizwan Khan66 and his group used poly(vinylpyrrolidone) (PVP) for the photocatalytic application of ZnO QDs (shown in Fig. 8a–e). They dissolved 0.3 g ZnAC in an ethanolic solution of PVP (0.05 g PVP in 50 mL ethanol). After 5 min stirring at 70 °C, they added 0.1 g NaOH for precipitation. This precipitate was then centrifuged at 3000 rpm for 5 min and washed with ethanol 3–4 times. Zwitterion-coated water stable ZnO QDs have also shown stability towards the dissolved salts. Zhang et al.79 have prepared zwitterion-coated ZnO QDs and investigated the stability of 2 mg mL−1 QDs over a period of 30 days. They used saturated sodium chloride solution for testing. They did not find any remarkable change in the absorbance for 30 days. The luminescence intensity decreased only 1% when the sample was kept at 4 °C but at 37 °C, the luminescence intensity decreased to 30% after 1 day and remained 13% after 30 days. Recently, an excellent application of SiO2-coated ZnO has been found for the fabrication of an ink-absorbing and UV-shielding film. This film is was made by the dispersion of SiO2-coated ZnO in the PVA matrix and then spin-coating this solution on to the PET substrate.
Fig. 8 (a) FESEM image, (b) TEM image of small ZnO QDs, (c) high-resolution TEM image of the ZnO QDs; the corresponding SAED pattern in the inset, (d) TEM image of the large ZnO QDs, (e) XRD pattern of the ZnO QDs.66 Adapted with permission from ref. 66. Copyright (2014) Elsevier and (f) TEM image of the PEG-capped ZnO QDs.65 Adapted with permission from ref. 65. Copyright (2017) Elsevier. |
The schematic of the whole process is shown in Fig. 9, schematic 1. The UV-visible absorbance spectra is shown in Fig. 9a–d to show the UV absorbance property of the film. Graphs were made for the various ratio of zinc acetate and NaOH and various reaction time as both these factors directly affect the size of the ZnO QD. The size difference ultimately affects the band gap of the particle. From this, we can clearly see that as the ratio increases, the absorbance peak shifts towards the lower wavelength side, showing a decrease in the size of the ZnO QDs.87
Fig. 9 Schematic 1 Synthesis of ZnO QDs@SiO2 and its addition into PVA for the fabrication of the ink-absorbing coating, (a–d) UV-visible absorbance spectra of the ZnO QDs prepared with different ratios of Zn(Ac)2-to-NaOH at 1:0.5, 1:1, 1:1.5, and 1:2, respectively.87 Adapted with permission from ref. 87. Copyright (2020) Royal Society of Chemistry. |
Weimin Yang and his group68 used the ultrasonication method and the micro-reactor method for the preparation of ZnO QDs; they named this method as the ultrasonic micro-reactor method (experimental setup is shown in Fig. 10a). In this method, two solutions, namely, zinc acetate and PEG-400 (n(PEG-400):n(Zn) = 1:1) in 50 mL ethyl alcohol and LiOH in ethyl alcohol, were injected through separate syringes into a PTFE tube, which was immersed in an ultrasonic washer. Oleic acid was used to precipitate the QDs, which were then collected by centrifugation. In this process, ultrasonication creates bubbles in the reaction solution due to ultrasonic cavitation. These bubbles are shown in Fig. 10b. These bubbles divide the reaction solution into several parts and restrict the growth of the QDs on the surface of the bubbles as the tensile stress is more on the surface of the bubbles. The effect of the flow rate on the size of QDs is found to be negligible as both the QDs, prepared at 300 μL min−1 and 750 μL min−1 at 40 °C and 180 W power, have nearly similar average size (shown in Fig. 10c and d). They thoroughly studied the effect of ultrasonic power, flow rate, and temperature on the synthesis and optical properties of the QDs. Fig. 10g shows the effect of reaction temperature on the emission and excitation wavelength. The corresponding graph of photoluminescence emission and excitation are shown in Fig. 10e and f, respectively. These two graphs are in close relation with Fig. 10g. Fig. 10h shows the variation in the quantum yield with temperature. Fig. 10k shows the effect of ultrasonic power on the emission and excitation wavelengths. The corresponding graph of photoluminescence emission and excitation are shown in Fig. 10i and j, respectively. Fig. 10l shows the variation in the quantum yield with ultrasonic power. Fig. 10o shows the effect of flow rate on the emission and excitation wavelength. The corresponding graphs of photoluminescence emission and excitation are shown in Fig. 10m and n, respectively. Fig. 10p shows the variation in the quantum yield with the flow rate. They successfully obtained a quantum yield of nearly 42%.68
Fig. 10 (a) Experimental setup of the ultrasonic microreactor, (b) simplified condition of the reaction solution in the tube under ultrasonication, (c and d) TEM and HRTEM micrographs, size distributions, electron diffraction patterns of the ZnO QDs synthesized under the flow rate of 300 lL min−1 and 750 lL min−1, respectively. The photoluminescence properties of ZnO QDs synthesized under the flow rate of 750 lL min−1 with 180 W ultrasonic power at different temperature. (e) Emission spectra, (f) excitation spectra, (g) emission and excitation peaks, and (h) quantum yield. Photoluminescence properties of the ZnO QDs synthesized at 40 °C under the flow rate of 300 lL min−1 with different ultrasonic power. (i) Emission spectra, (j) excitation spectra, (k) emission and excitation peaks, and (l) quantum yield. Photoluminescence properties of the ZnO QDs synthesized at 40 °C with 180 W ultrasonic power at different flow rates; (m) emission spectra, (n) excitation spectra, (o) emission and excitation peaks, and (p) quantum yield.68 Adapted with permission from ref. 68. Copyright (2016) Elsevier. |
Fig. 11 (a–d) Illustration of ZnO–GO formation, (e and f) PL spectra of the ZnO–GO composite and (g) the energy level of the ZnO–GO composite.16 Adapted with permission from ref. 16. Copyright (2020) American Chemical Society. |
Some researchers have fabricated the ZnO–silica composite using tetraethoxysilane (TEOS). Recently, Liang et al.76 have mixed different amounts of ZnO QDs in the aqueous solution of TEOS. Patra and his group67 have also prepared ZnO QDs using TEOS. In a simple sol–gel method, first, they prepared an ethanolic solution of ZnAc. Then, they added KOH for maintaining the pH of the above solution at 10, 12, and 14. Then, they added TEOS to the above solution. The obtained colloidal solution was centrifuged and washed several times with methanol and water. Finally, the colloid was dispersed in water. Zain et al.78 also used TEOS for the synthesis of ZnO QD-embedded silica nanoparticles. They found 55% to 80% luminescence emission at 100 °C temperature and 40 g L−1 salinity. This increase in the stability was due to the hydrophobic properties of silica. You Liang and his group77 have prepared kasugamycin (KAS)-conjugated ZnO QDs for a pH-responsive pesticide delivery system. It was found to be more effective against bacterial fruit blotch as compared to pure KAS or ammonia-treated ZnO QDs.
For the synthesis of KAS-conjugated ZnO QDs, the QDs were prepared first, then (3-aminopropyl)trimethoxysilane (APTES) was modified on the surface of the QDs. The complete modification process is composed of three steps, which is shown in Fig. 12. From this figure, we can see that APTS-conjugated ZnO QDs then react with 4-formylbenzoic acid, followed by KAS to form the KAS-ZnO QDs.
Fig. 12 Synthesis mechanism for KAS-conjugated ZnO QDs.77 Adapted with permission from ref. 77. Copyright (2018) Elsevier. |
In summary, there is only one method based on the precipitation of zinc salt (mostly zinc acetate) in alcoholic medium. Different researchers have used different precipitating agents such as KOH, NaOH, and LiOH. A variation in precipitation results in a variation in the different physical properties. It was also found that at different pH values, we also get a variation in the band gap. Beside simple precipitating methods, there are some methods that are based on the ultrasonication of the precursor solution. Some methods have also been developed for more stable QDs, which are based on different capping agents such as SiO2, oleic acid, NIPAM, TBAM, APM, PEG, and PVP. The capping results are enthusiastic and produce more stable QDs as compared to pristine ZnO QDs. Conjugating organic compounds with QDs also results in stable QDs. Composites made in such a form are useful in a number of applications. The composites can be stable in salt water, which makes QDs applicable in saline conditions. Composites made up of graphene have tremendous potential in the near future (Fig. 13).
Fig. 13 (a) Experimental set-up of the RF plasma reactor.55 Adapted with permission from ref. 55. Copyright (2020) IOP Science and (b) plasma reactor chamber.95 Adapted with permission from ref. 95. Copyright (2014) Wiley. |
Fig. 14 (a) XPS spectra of Pr3+-doped ZnO QDs, (b) binding energy scan of Pr3+-doped ZnO QDs for Pr 3d5/2, (c and d) fitting spectra of O 1s from undoped and 2 mol% Pr3+-doped ZnO QDs, respectively and (e) XPS spectrum of Pr 3d5/2 from 2 mol% Pr3+-doped ZnO QDs.38 Adapted with permission from ref. 38. Copyright (2014) Springer. |
This result was due to the variation in the chemical bonding environment of Pr3+, which confirms the doping of Pr3+ in the ZnO lattice. From the XPS spectra of the Pr 3d5/2 peak (shown in Fig. 14e), we can observe an increase in the intensity with the increase in the doping concentration of Pr3+, which shows that the Pr3+ concentration increases in the ZnO QD with concentration. The spectrum of Pr 3d5/2 from 2 mol% Pr3+-doped ZnO QDs shows two peaks at 934.0 and 931.1 eV (shown in Fig. 14c). The higher binding energy component (HBEC) of Pr 3d5/2 at 934.0 eV is due to the formation of the Pr–O–Zn bond and the lower binding energy component (LBEC) of Pr 3d5/2 at 933.1 eV is due to the formation of the Pr–O–Zn bond in the electron rich-environment formed by oxygen vacancies. The O 1s spectra also have HBEC and LBEC similar to Pr 3d5/2 (shown in Fig. 14c and d). The HBEC of O 1s at 532.0 and 531.9 eV is due to the O–Zn bond surrounded by electron vacancies. The LBEC of O 1s at 530.5 eV is due to oxygen in the O–Zn bond. Both these spectra show a significant increase in the FWHM of all the components of LBEC and HBEC. This broadening is due to the increase in the distortion of the electron cloud of oxygen in the Pr3+-doped samples. As most of the doping elements are doped at the surface of the ZnO crystal and the XPS studies are also limited to a few nm of the crystal surface, it is an effective method for the confirmation of doping. From the XPS full spectrum scanning,36 we can see that the diffraction peaks of zinc and oxygen are only present before the doping of La. Sun et al.35 found that the size of ZnO QDs decreases with the increase in the La content.
Jakub Sowik84 and his group extensively studied the effect of different rare earth elements in different proportions on the optical, structural, and photocatalytic properties of ZnO QDs. The PL spectra of different rare earth metal-doped ZnO QDs is given in Fig. 15a and b. Here also, the XPS spectra of different rare earth elements show the binding energy shift in Zn 3d and the valence band spectra, which confirms the doping of rare earth elements. Er-Doped ZnO QDs was found to be the most effective towards the decomposition of phenol solution under visible light. The activity toward phenol decomposition was nearly 90% in UV. The PL intensity of La-doped ZnO QDs shows the highest PL quantum yield (nearly 80%). Gd doping is not only found to increase the quantum yield from 31% to 94% but also produces a red shift in the absorption and emission spectra.99 Similarly Huang et al.36 synthesized La-doped ZnO QDs and studied the effect of the OH− ion and La concentration on the oxygen vacancy defect, which ultimately affects the fluorescence performance of the ZnO QDs. The oxygen vacancy was maximum for the 1:1 OH−:Zn molar ratio and 7 mol% doping amount of La. The fluorescence emission intensity was found to be in direct proportion to the oxygen vacancy concentration. The fluorescence stability for 5 mol% doped QDs was nearly 5 months (shown in Fig. 15c). Rare earth doping will be very helpful in increasing the quantum efficiency of solar cells via down-conversion. Trivalent praseodymium could be an encouraging material as an activator for the host material. It was found that Pr3+doping effectively increases the PL intensity of some (578, 630, and 752 nm) defect-related emissions in the ZnO QDs.32 This increase in the emission is mainly related to energy transfer from the Pr3+ ion to the defect states.
Fig. 15 (a and b) PL spectra of different rare earth metal-doped ZnO QDs.84 Adapted with permission from ref. 84. Copyright (2018) Elsevier. (c) 5 mol% La-doped ZnO QDs at different intervals of time and (d) Pr3+-doped ZnO QDs.32 Adapted with permission from ref. 32. Copyright (2014) Elsevier. |
From Fig. 15d, we can see that with the increase in the Pr concentration, the PL intensity also increases up to 1% doping concentration and then decreases for 2%. Thus, we can say that 1% is the doping limit for fluorescence applications. This decrease in the PL intensity could be a result of the concentration quenching effect. This effect is a result of cross relaxation and energy migration when the concentration of doping elements exceeds certain limits.
Fig. 16 (a) XRD pattern of Zn1−xGdxO, the TEM images of the Zn1−xGdxO QDs at (b) x = 0, (c) x = 0.01, (d) x = 0.03, and (e) x = 0.07 (inset HRTEM images). The plot of (f) band gap, (g) Stokes shift, (h) PL intensity, and (i) FWHM in terms of the amount of Gd ions. (j) Quantum yield and (k) charge transfer mechanism.99 Adapted with permission from ref. 99. Copyright (2019) Elsevier. |
Besides, rare earth transition metal-doped ZnO QDs have been found to show dilute magnetic properties.103 As a consequence of their tremendous application in spin-electronic and spin-photonic devices, vigorous investigation has to be done in the field of ferromagnetism in QDs. Magnetism in semiconductors arises due to the magnetic exchange interactions between the delocalized charge carriers and the localized magnetic impurities. These interactions result in large Zeeman and Faraday rotation effects. Mostly, in the early years, ligand-to-metal charge transfer(LMCT)104,105 transition was found to be the cause of ferromagnetism in doped ZnO QDs.83 In nickel and cobalt-doped ZnO QDs, the LMCT transition was also found to be responsible for ferromagnetism.106 This was also found to show a large Zeeman effect.107,108 In both these cases, there was a sub band gap in the energy level due to this LMCT. Beside this sub band gap, unique midgap excited states have been found by Joseph W. May in cobalt-doped ZnO QDs.109 Midgap transitions give rise to a number of properties such as magneto-electronic, magneto-optic, photocatalytic, and sensing properties. Stefan et al.110 later found a colloidal analogue to bind the magnetic polaron (BMP)111 responsible for magnetism in Mn2+-doped ZnO QDs. Yong112 and his group, via DFT theory,85,113 found that in the case of Mn2+-doped ZnO QDs, magnetism rises due to double exchange in the charge-transfer excited states. In case of p-type (N2− doped) Mn-doped ZnO QDs, magnetic interaction between two Mn2+ ions is mediated by the N2− ion. This results in parallel alignment of the spin of both the Mn2+ ion, which gives rise to ferromagnetism at room temperature.114
Transition metal doping suppresses the growth of ZnO QDs and increases the various vacancy defects. Zhang et al.33 have also found similar results for Cd-doped ZnO QDs. With the increase in the Cd concentration, there was a blue shift in the UV spectra due to the quantum confinement effect. From Fig. 17a, a blue shift in the exciton absorption spectra was observed and from Fig. 17b, we can also see that there was a shift in the direct band gap for different concentrations of Cd. The PL intensity for the emission spectra was also found to increase with the increase in the doping concentration. Similarly, copper doping was also found to alter the energy band gap of the ZnO QDs. The energy band gap decreases with the concentration of Cu. From Fig. 17c, we can see that absorption increases with the doping concentration, which is mainly because of the substitution of Zn2+ by Cu2+. The substitution of Zn2+ results in an increase in the oxygen vacancies and electron concentration.39
Fig. 17 (a) UV-visible absorbance spectra of Cd-doped ZnO QDs at different concentration of Cd, (b) plots of (αhν)2 versus hν of ZnO QDs with different concentrations of Cd.33 Adapted with permission from ref. 33. Copyright (2012) Springer. (c) UV-visible absorbance spectra of Cu-doped ZnO QDs.39 Adapted with permission from ref. 39. Copyright (2016) Elsevier. (d–f) UV-vis absorption spectra, plot of (αhν)2 vs. (hν) and calculated Egap of Sn2+-doped ZnO QDs, respectively. (g–i) UV-vis absorption spectra, plot of (αhν)2 vs. (hν) and calculated Egap of Sn4+-doped ZnO QDs, respectively.85 Adapted with permission from ref. 85. Copyright (2017) Royal Society of Chemistry. |
With Sn doping, it was found that when the doping concentration of Sn was less than 3%, the doping was interstitial, and when the doping concentration was greater than 5%, the doping was substitutional.85 The UV-visible spectrum of Sn2+-doped ZnO QDs is shown in Fig. 9d and the corresponding change in the energy band gap is shown in Fig. 17e and f. From Fig. 17f, we can clearly see that with the increase in the Sn2+ concentration, initially, the band gap energy increases from 3.542 eV to 3.572 eV, then drops to 3.418 eV, and is finally increased to 3.480 eV. Similar results were shown by Sn4+-doped samples (shown in Fig. 17g–i). In case of the Sn4+-doped sample, the energy band gap first increased from 3.542 eV to 3.613 eV, then dropped to 3.510 eV, and finally rose to 3.530 eV. The Egap reached the minimum value when the Sn4+ concentration was 0.05. This not only affects the binding energy of the O 1s spectrum but also affects the concentration of different types of defects present in the QDs.
Fig. 18 TEM images and SAED patterns of (a) undoped ZnO and CexMg0.1Zn0.9−xO QDs with (b) x = 0, (c) x = 0.004, (d) x = 0.008, and (e) x = 0.01.115 Adapted with permission from ref. 115. Copyright (2020) Elsevier. |
Fig. 19 (a) EPR spectra of undoped ZnO and CexMg0.1Zn0.9−xO QDs, (b) EPR and PL (inset) spectra of CexMg0.1Zn0.9−xO QDs, O 1s XPS spectra of CexMg0.1Zn0.9−xO QDs with (c) x = 0.004, (d) x = 0.008, (e) x = 0.01, and (f) Ce 3d XPS spectra of CexMg0.1Zn0.9−xO QDs (x = 0.004, 0.008, 0.01).115 Adapted with permission from ref. 115. Copyright (2020) Elsevier. |
In summary, doping in QDs is not as easy as in nanoparticles. Doping mainly results in the smaller size of QDs as it restricts the particle growth. Doping also alters the energy band gap of the QDs, which causes a blue shift. Larger doping elements mostly reside on the surface of the QDs. But there are some methods that help the doping elements to reside at the core of the QDs. Rare earth doping mostly results in the manipulation of the luminescence properties. It is also possible to use these as photocatalysts. Besides, rare-earth doping and transition metal doping is found to show magnetism in the QDs. Here, doping elements not only reside on the surface but also inside the core up to some limit. Magnetic interactions between the spins are mostly LMCT and double exchange interactions. These result in room temperature ferromagnetism as well as large Zeeman and Faraday rotation effects.
Fig. 20 (a) F0/F for different ions, (b) fluorescence intensity of the Fe2+ ion at different concentrations.40 Adapted with permission from ref. 40. Copyright (2019) American Chemical Society. (c) F0/F for different ions, (d) fluorescence intensity of the Cr6+ ion at different concentrations.37 Adapted with permission from ref. 37. Copyright (2019) IOP Science. (e) Fluorescence intensity for different ions, (f) bar diagram for different ions.146 Adapted with permission from ref. 146. Copyright (2019) Elsevier. (g) F0/F for different ions and (h) fluorescence intensity of the chlorine ion at different concentrations.147 Adapted with permission from ref. 147. Copyright (2016) Royal Society of Chemistry. |
As the concentration of chlorine increases, it absorbs electrons from the oxygen vacancies. This results in a decline in the emission intensity of the QDs. From Fig. 20g, we can see the F0/F for different ions, which shows selectivity towards chlorine in the solution. Fig. 20h shows the fluorescence spectra for the chlorine ion at different concentrations. Capping with different polymers or organic materials produces selectivity toward different polluting elements. These ZnO QD-based detecting solutions are very cheap and easy to handle. Thus, these QDs can become a useful tool for the detection of impurities (Fig. 21).
Fig. 21 Gas response curve at (a) different operating voltages and H2 concentrations, the response and recovery curve in the inset, (b) different gases.148 Reprinted (adapted) with permission from ref. 148. Copyright (2019) IOP Science. Response curve for (c) different NO2 concentrations and (d) different gases.149 Adapted with permission from ref. 149. Copyright (2011) Royal Society of Chemistry. (e) Real time response curve and (f) ratiometric calibration curve for CaDPA detection.151 Adapted with permission from ref. 151. Copyright (2017) Royal Society of Chemistry. |
The sensing mechanism of the QD-based sensors is quite different as compared to nanoparticle-based sensors. The change in the electrical conductivity is the basis of the sensing property of the nanoparticles.152–154 The red-emitting Eu(III) ion acts as a signal reporting unit via chelation with calcium dipicolinate (CaDPA), which is the biomarker of Bacillus anthracis spores.
Chen et al.156 have used QDs as a biosensor for the detection of histone acetylation using acetyl coenzyme A (Ac-CoA) as the target molecule. This is an indirect method for the detection of histone acetyltransferase (HAT). The schematic diagram for the detection mechanism is given in Fig. 22a and b. In this process, CoA is produced as a by-product of acetylation of Ac-CoA, which can be easily detected by the photo-electrochemical biosensor due to the presence of phosphate and thiol groups in its structure. From Fig. 22c, we can see that with the increase in the HAT concentration, the photocurrent also increases gradually. Fig. 22d shows a linear relation between the logarithmic concentration of HAT and the photocurrent intensity. From its wide range of linear relation, we can predict its potential for the detection of HAT. It shows high selectivity towards the HAT molecule in comparison to other molecules (as shown in Fig. 22e). It has relatively similar cycles for detection (as shown in Fig. 22f), having nearly 0.92% calculated relative standard deviation (RSD). Suppressing the working of electroluminescence reagents is emerging as an effective tool for the detection of biomolecules.157 β-Cyclodextrin (β-CD)-capped ZnO QDs decorated with pyridoxal 5′-phosphate (PLP) and pyridoxal (Py) is useful for the detection of histamine.158
Fig. 22 (A) Mechanism for the acetylation of the short peptide catalyzed by HAT, (B) schematic diagram of the construction of the PEC biosensor for detecting HAT, (C) the photo-electrochemical response of the biosensor with different concentrations of HAT. (D) Calibration curve of the biosensor for HAT detection. (E) The histogram for the photocurrent changes of the biosensor fabricated with different targets. The concentration of different targets was 100 nM. (F) Time-based photocurrent response of the biosensor toward 100 nM HAT.156 Adapted with permission from ref. 156. Copyright (2020) Elsevier. |
Fig. 23 (a–d) Illustration of the ZnO/HPEI nanocomposite and the fluorescent cell imaging, UV-Vis and PL spectra of ZnO QDs at [Zn2+/OH−] molar ratio of (e and f) 1:2, (g and h) 1:1, and (i and j) 1:4.124 Adapted with permission from ref. 124. Copyright (2020) Elsevier. |
Similarly, Liu et al.123 have used ZnO QDs for the bio-imaging of HeLa cells. They synthesized the SiO2-coated ZnO QDs. For chemo-luminescence bio-imaging, ZnO/SiO2 was dissolved in ultrapure water and then it was used for staining the culture-grown HeLa cells. These samples were then added to CPPO, which was then cross-linked with F127 for improving the hydrophobicity of CPPO. After adding H2O2, this sample was ready for imaging. This whole process is shown in Fig. 24a.
Fig. 24 (a) Schematic illustration of ZnO NPs@SiO2-based CL bio-imaging. (b and c) CL images (b) and intensities (c) of the HeLa cells cultured with ZnO@SiO2 NPs for 6 h and then added into the mixture of CPPO at different concentrations of H2O2. (d) Cell viability of the HeLa cells after 24 h incubation in different concentrations of the ZnO@SiO2 NPs.123 Adapted with permission from ref. 123. Copyright (2020) Elsevier. |
With increasing H2O2 concentration, the intensity of imaging was also found to be increase (shown in Fig. 24b and c), which indicates a direct relation between H2O2 and the imaging intensity. However, the cell viability for different concentrations of ZnO/SiO2 after 24 h of incubation remains nearly constant (shown in Fig. 24d).
The electron transport layer (ETL) plays a major role in any optoelectronic device. It should be capable of quickly transporting electrons so as to avoid charge recombination. Having high electron mobility, ZnO plays this role very effectively in most optoelectronic devices. Fig. 25g and h shows the schematic diagram of a ZnO–GO-based QLED device and the energy level diagram, respectively. Fig. 25i shows the blue electroluminescence (EL) of the device. Fig. 25k shows an increase in the luminance and the current density with voltage. Fig. 25l shows the nearly constant luminous efficiency and the quantum efficiency with increasing voltage. Fig. 25m shows the practical 100 pixels 8 V display of the blue QLED. It has been found that the ETL (electron transport layer) made of ZnO QDs has more current efficiency in comparison to organic ETL. The energy level diagram of the ZnO QD-ETL based device has been shown in Fig. 25a, which shows that the electron can be easily transported through the ETL to the Al cathode. It also helps in blocking the hole movements. Fig. 25e and f shows the device structure of the QLED-based on organic ETL and ZnO ETL. It was due to better charge balance by the ZnO QDs.164 The efficiency can be further increased with the help of doped ZnO QD-ETL.165 Doping in the ETL lifts the conduction band minima and reduces the electron mobility. The energy level diagram (Fig. 25b) shows a shift in the conduction band minima with the doping concentration of Mg. The TRPL spectra of the device (shown in Fig. 25c) shows a reduction in exciton quenching at the interface. This is very essential for improving the device performance. Fig. 25d shows a reduction in the current density with increasing Mg concentration. This is helpful in increasing the charge balance. Doping also results in an increase in the external quantum efficiency. Shi et al.115 fabricated Mg- and Ce-doped ZnO QDs LED for white light emission. The EL spectrum of the device is shown in Fig. 25e and the device image is shown in the inset. At 3 V potential and 200 mA driven current, this LED-correlated color temperature (CCT) and the color rendering index (CRI) were 5733 K and 81, respectively. For any optoelectronic device, ETL is an important part.166 Because of its low fabrication temperature and high electron mobility in comparison to TiO2,167 ZnO is one of the best candidates for ETL. Even in perovskite-based devices, ZnO is preferred as the ETL. In perovskite-based solar cells, ZnO increases the Voc, which results in an increase in the fill factor and efficiency.167 The high electron mobility of ZnO is useful in high electron extraction. It improves the band alignment with inorganic perovskite materials, improves the device stability, and reduces interfacial non-radiative recombination.168 The introduction of ZnO in inorganic perovskite films results in more compactly and uniformly distributed crystalline grains as it is evident that perfectly oriented grains considerably increase the power conversion efficiency.169 The ETL of ZnO in this case also enhances the transport of photo-generated carriers from the perovskite film to the electrodes, which improves the rise and fall time of perovskite-based photodetectors.170
Fig. 25 (a) Energy level of QLED.164 Adapted with permission from ref. 164. Copyright (2014) American Scientific. (b) Energy level of Mg-doped ZnO-based QLED, (c) TRPL spectra of Mg-doped ZnO and the QDs interface, (d) J–V curve of the Mg-doped ZnO device.165 Adapted with permission from ref. 165. Copyright (2020) Elsevier. The device structure of (e) 3TPYMB and (f) ZnO QDs.164 Adapted with permission from ref. 164. Copyright (2014) American Scientific. (g) ZnO–GO based QLED, (h) energy level diagram, (i) EL spectra of the ZnO–GO-based device.16 Adapted with permission from ref. 16. Copyright (2020) American Chemical Society. (j) EL spectra of Ce- and Mg-doped ZnO QLEDs.115 Adapted with permission from ref. 115. Copyright (2020) Elsevier. (k) Current density and luminance vs. voltage. (l) Luminous efficacy and external quantum efficiency (EQE) vs. voltage and (m) 100 pixel display of the ZnO–GO-based QLED.16 Adapted with permission from ref. 16. Copyright (2020) American Chemical Society. |
Transparent nanocomposites made from ZnO QDs, SiO2, and epoxy171 for encapsulating LED172 would be highly beneficial in the near future. Besides encapsulating LED, nanocomposites are also applicable in other fields. Li et al.173 prepared ZnO QD-based polymer composite using poly(methyl methacrylate) (PMMA) as the polymer for UV-shielding87 material. This material can be used in a number of applications such as contact lenses, UV-shielding windows, or glasses. It was also found that the ZnO QDs can initiate photo-degradation of the host polymer matrix, which can be applied for recording materials. Georgia G. Goourey81 and his team studied the effect of ZnO nanoparticles with size ranging from 5 to 30 nm on acrylate photopolymers. They found partial quenching of the yellow green fluorescence with the incorporation of the QDs, whereas with the increase in the size to 30 nm, photo-degradation decreases drastically. ZnO QDs were also found to degrade organic compounds, which are poisonous to human health. Fakhri et al.82 prepared the ZnO QD/CuO composite for the degradation of the Tetanus toxin. They found 75% degradation under UV irradiation.
Fig. 26 (a) Schematic diagram for charge generation on the surface of the ZnO QDs and the energy level diagram, absorption spectra of the dye for the decolorization process with initial diameter (b) 3.2 nm, (c) 5.8 nm, (d) 6.7 nm, and band gap and size of the growing ZnO QDs for particles with starting diameters of (e) 3.2 nm, (f) 5.8 nm, (g) 6.7 nm.24 Adapted with permission from ref. 24. Copyright (2020) American Chemical Society. |
From Fig. 27, we can see the whole killing mechanism. In detail, the anionic membrane is first attacked by the cationic ZnO QDs composite, which destroys the cell membrane of the bacteria. When the composite attacks the membrane, it releases Zn2+ ions due to the lower pH level at that region. The absorption of Zn2+ on the surface of the membrane not only inhibits the respiratory action of the enzymes but also produces reactive oxygen species (ROS). ROS damages the membrane, DNA, and mitochondria. GO sheets produce hyperthermia, which prevents infections in the wound. Abinit Saha and his group174 found that QDs having particle size in the range of 3–5 nm have the maximum ability to destabilize the CRP (Cyclic AMP Receptor Protein) structure of the E. coli bacteria.
Fig. 27 Schematic diagram for the killing mechanism of Staphylococcus aureus and Escherichia coli bacteria.43 Adapted with permission from ref. 43. Copyright (2019) Elsevier. |
Fig. 28 Schematic illustration of the synthesis and working process of folic acid-conjugated ZnO QDs.42 Adapted with permission from ref. 42. Copyright (2011) Royal Society of Chemistry. |
Lastly, we can say that ZnO can help in cancer treatment not only by its killing ability but also by acting as a drug carrier. It could play a major role in overcoming the challenges such as recurrence of disease, drug resistance, and side-effect of the drug. Drug resistance is found with many known anti-cancer drugs currently in use. Recurrence poses a challenge that is also related to these known drugs. ZnO will be very helpful in overcoming these challenges. Its drug delivery mechanism will improve the selectivity of the drug towards cancer cells, which finally increases its effectiveness. It will also reduce the possibility of drug side-effects as it will only act at the specific site.
S. no. | Sample name | Reversible capacity (mA h g−1) | Cycle life | Capacity retention (%) | Current rate (mA g−1) | Ref. |
---|---|---|---|---|---|---|
1 | Amorphous ZnOQDs/MPCBs | 930 | 85 | 90 | 100 | 176 |
840 | 280 | 93.1 | 200 | |||
510 | 400 | 94 | 1000 | |||
2 | ZnO-VAGNs | 809 | 100 | 93 | 80 | 177 |
450 | 250 | 87.7 | 350 | |||
3 | ZnO-QDs@CMS | 1015 | 80 | — | 50 | 178 |
565 | 350 | 94.3 | 1000 | |||
4 | ZnO@ZnO QDs/C NRA | 699 | 100 | 100 | 500 | 179 |
5 | ZnO QDs@porous carbon | 1150 | 50 | 67.6 | 75 | 180 |
6 | ALD ZnO/G | ∼540 | 100 | 100 | 181 | |
7 | ZnO/RGO | ∼800 | 200 | 104 | 200 | 117 |
668 | 700 | 87 | 1000 |
Initial cathodic scanning shows a peak at 0.31 V, which is attributed to the reduction of Zn2+ to Zn metal and the generation of a solid electrolyte interphase. The peaks at 0.69 and 0.37 V are due to the reduction of ZnO to Zn and the formation of the LixZn alloy with Li+, respectively. The de-alloying of LixZn is clearly visible from the four peaks of the anodic scanning curve at 0.27, 0.36, 0.55, and 0.68 V. A large peak at 1.3 V in the anodic scanning curve is due to the formation of ZnO in the reaction of Zn and Li2O (shown in Fig. 29a). From Fig. 29b, we can see that the first charging and discharging curve shows a voltage at 1.3 V and a large plateau at 0.3 V. The initial discharge and charge capacity were 1027 and 766 mA h g−1, respectively, which gives an initial coulombic efficiency of 74.6%. The cyclic performance of the ZnO/RGO composite electrode and ZnO@RGO at a current density of 200 mA g−1 gives 766 mA h g−1 initial reversible capacity, which drops to 605 after 20 cycles (shown in Fig. 29c). After 30 cycles, the coulombic efficiency stabilizes above 97% and after 200 cycles, the reversible capacity reaches 800 mA h g−1. Fig. 29d shows the average reversible capacities of 840, 670, 590, 515, 400, and 315 mA h g−1 at 0.1, 0.5, 1.0, 2.0, 5.0, and 10.0 A g−1, respectively, for the ZnO/RGO composite electrode. Small ZnO size and the large and thin supporting structure of RGO are responsible for the excellent rate capability of the ZnO/RGO electrode. The cyclic performance of the ZnO/RGO electrode at a higher current density of 1000 mA g−1 is shown in Fig. 29e.
Fig. 29 (a) CV curve for 5 initial cycles, (b) charge–discharge curve, (c) cycling performance, (d) rate performance, and (e) cycling performance at 1000 mA g−1 for the ZnO/RGO composite electrode.117 Adapted with permission from ref. 117. Copyright (2020) Royal Society of Chemistry. |
In summary, the main applications of doped and undoped ZnO QDs are found to in antibacterial, antifungal, and anticancer medicines. Other applications are the sensing and detection of toxic materials. Quantum confinement enhances the properties of the ZnO QDs, which opens the door for noble applications. pH-responsive drug delivery will be game changing in the treatment of cancer. The photocatalytic activity will revolutionize its application in different chemical industries and environment-related challenges. Graphene and polymer-based composites can widen the application field of ZnO QDs. Graphene and reduced graphene-based ZnO composites are better candidates for storage applications in the future.
(1) With a reduction in the particle size below 8 nm, strong quantum confinement effect comes into play. Due to exciton–phonon interaction, the ground state energy and oscillator strength reduces with the particle size of the QDs. Polaronic self-energy corrections of the exciton vanishes completely and the PB potential effectually transforms into a dynamically-screened Coulomb potential.
(2) Longitudinal optical (LO) and transverse optical (TO) modes of bulk ZnO splits into A1 and E1 symmetries in ZnO QDs. The phonon peak shift that arises in the ZnO QDs is related to three main factors: phonon localization by defect creation, confinement effect within the QD boundaries, and the laser-induced heating effect.
(3) The fluorescence intensity of the QDs increases with the doping of elements. But after a certain limit, it starts decreasing. The fluorescence intensity is quenched with the addition of any metal ion impurity or organic ligand. This property is useful for fluorescence probes. The surface charge density also increases in the QDs.
(4) The sol–gel method is the most popular as it is a relatively more efficient, simple, and inexpensive method over others. With the help of this method, QDs stable in water and other alcohols (mostly ethanol) can be prepared. Composites made via SiO2, graphene, and PMMA are able to increase the stability of the QDs. Radio frequency-based synthesis is emerging as a new technique for QD synthesis.
(5) Most of the doping elements are found on the surface of the QDs. With the help of the core/shell (ICS) procedure, we can inject doping elements inside the core of the ZnO QDs. The doping of rare earth elements is difficult as compared to transition metal doping. Rare earth elements are useful for optical properties and transition metal doping induces magnetism in the QDs.
(6) Rare earth doping is mostly favorable for the variation in the photoluminescence-related properties, whereas transition metal doping mostly induces magnetic behavior in the QDs. Both types of doping introduces photocatalytic character in the QDs by producing different vacancies.
(7) The detection of metal ion impurities (Fe2+, Cr6+, Hg2+, Cl−, etc.) distinguishes it from nanoparticles. The interaction of graphene and QDs is helpful in optoelectronic device fabrication. The enhancement in ROS generation increases its anti-bacterial and anti-microbial activity. ROS generation also increases the cytotoxicity of the QDs. An effective drug delivery mechanism sharpens the QDs' cytotoxicity. Li-Ion storage can be achieved by hanging QDs in the framework of graphene or graphene-derived components.
(8) Having such a wide field of application, ZnO QDs have a very bright future. They can be used as an impurity-detecting tool. They can be used for the production of low-cost gas sensors. They can replace TiO2 from the ETL in many optoelectronic devices. Apart from their commercial use in cosmetics, they can also be used in anti-microbial and anti-bacterial ointments. Their use in the treatment of cancer will definitely increase the effectiveness of the drug and will decrease the side-effects of the drug at the same time.
This journal is © The Royal Society of Chemistry 2021 |