Next Article in Journal
Study on the Adsorption Properties of Graphene Oxide/Laponite RD/Chitosan Composites
Previous Article in Journal
Structure, Morphology, Heat Capacity, and Electrical Transport Properties of Ti3(Al,Si)C2 Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Enhanced Optical and Antibacterial Activity of Hydrothermally Synthesized Cobalt-Doped Zinc Oxide Cylindrical Microcrystals

by
Awais Khalid
1,*,
Pervaiz Ahmad
2,*,
Abdulrahman I. Alharthi
3,
Saleh Muhammad
1,
Mayeen Uddin Khandaker
4,
Mohammad Rashed Iqbal Faruque
5,
Abdulhameed Khan
6,
Israf Ud Din
3,
Mshari A. Alotaibi
3,
Khalid Alzimami
7,
Abdulrahman A. Alfuraih
7 and
David A. Bradley
4,8
1
Department of Physics, Hazara University Mansehra, Khyber Pakhtunkhwa 21300, Pakistan
2
Department of Physics, University of Azad Jammu and Kashmir, Muzaffarabad 13100, Pakistan
3
Department of Chemistry, College of Science and Humanities, Prince Sattam Bin Abdulaziz University, P.O. Box 173, Al-Kharj 11942, Saudi Arabia
4
Center for Applied Physics and Radiation Technologies, School of Engineering and Technology, Sunway University, Bandar Sunway 47500, Selangor, Malaysia
5
Space Science Centre, Universiti Kebangsaan Malaysia (UKM), Bangi 43600, Selangor, Malaysia
6
Department of Biotechnology, University of Azad Jammu and Kashmir, Muzaffarabad 13100, Pakistan
7
Department of Radiological Sciences, College of Applied Medical Sciences, King Saud University, P.O. Box 10219, Riyadh 11433, Saudi Arabia
8
Department of Physics, University of Surrey, Guilford GU2 7XH, UK
*
Authors to whom correspondence should be addressed.
Submission received: 3 May 2021 / Revised: 4 June 2021 / Accepted: 7 June 2021 / Published: 11 June 2021

Abstract

:
Cobalt (Co) doped zinc oxide (ZnO) microcrystals (MCs) are prepared by using the hydrothermal method from the precursor’s mixture of zinc chloride (ZnCl2), cobalt-II chloride hexahydrate (CoCl2·6H2O), and potassium hydroxide (KOH). The smooth round cylindrical morphologies of the synthesized microcrystals of Co-doped ZnO show an increase in absorption with the cobalt doping. The antibacterial activity of the as-obtained Co-doped ZnO-MCs was tested against the bacterial strains of gram-negative (Escherichia coli, Klebsiella pneumonia) and gram-positive bacteria (Staphylococcus aureus, Streptococcus pyogenes) via the agar well diffusion method. The zones of inhibition (ZOI) for Co-doped ZnO-MCs against E. coli and K. pneumoniae were found to be 17 and 19 mm, and 15 and 16 mm against S. Aureus and S. pyogenes, respectively. The prepared Co-doped ZnO-MCs were thus established as a probable antibacterial agent against gram-negative bacterial strains.

1. Introduction

The difference of dimensions seen between the atomic and molecular scale of basic science and the microstructural scale of engineering and fabrication is balanced by the characteristic function of nanoscale materials [1]. Scientific research on crystalline nanomaterials has evolved exponentially at various levels over the last few decades. Several visualized application possibilities for such novel materials encourage intensive investigations. Semiconductors of the group (II–VI) are commonly studied because of their novel size-dependent electrical, optical, and optoelectronic properties.
Zinc oxide (ZnO) is among the most promising semiconductor materials from many perspectives [2,3,4]. It is an eco-friendly, thermally stable, biocompatible, versatile material with the potential to experience photocatalysis in a neutral, basic, and acidic medium. Such attractive properties, easy method of synthesis, and sophisticated growth develop ZnO-based devices in photonics, electronics, sensing, and acoustics [5,6]. This material is promising for optoelectronic devices of short wavelength [7,8], due to very high exciton binding energy (60 MeV), high photosensitivity, high electron mobility, inexpensive route of synthesis with diverse morphologies, flexibility in chemical functionality, biocompatibility, high transparency, and wide direct band gap (Eg = 3.37 eV) at room temperature [8]. Due to the huge electronegative value difference between O2− (3.44) and Zn2+ (1.65), the bonding in ZnO is ionic. Even so, the alternating layers populated by Zn2+ and O2− atoms form a crystal structure in which Zn2+ (cation) is coordinated tetrahedrally with four O2− (anions). The non-centrosymmetric structure that results from this coordination (tetrahedral) produces pyroelectric and piezoelectric properties of ZnO [9]. The effective way of getting unique properties in ZnO is the introduction of impurities into the ZnO host. In the past decades, several research groups have done interesting research to work widely on the unique arrangement of transition metal (TM) ion-doped ZnO nanoparticles with magnetic and optical properties [10,11]. The TM oxide-doped ZnO holds place in various applications including sensors [12], solar cell [13], optoelectronics [14], spintronics [15], and piezoelectric devices [16]. The analysis of the variations in the properties arising due to doping of different transition metals such as Cu, Ni, Co, Mn, Cr, and Fe to ZnO has always been the matter of controversial studies [17,18,19,20,21,22,23,24,25].
Cobalt has its significance among the various TM ions due to its comparable ionic radius (0.74 Å) to that of ZnO (0.745 Å) [26]. It can change the morphology and properties of ZnO nanostructures. It has a strong magnetic moment compared to other transition metals [2]. Co-doped ZnO diluted magnetic semiconductor (DMS) may possess the ferromagnetic property at room temperature [27]. Similarly, Co-doped ZnO nanostructures are found to have more thermal stability in comparison to pure ZnO [28,29]. Co-doped ZnO structures have many potential applications in medical [30], electronics [31], photo-catalysis [32], solar cells [33], thermoelectric [34], 3D printing [35], light-emitting diodes [36], humidity sensors [37], and biosensors [38]. It can also be used as an antifungal and antibacterial agent [39]. Co-doping also makes ZnO more flexible for the above-mentioned applications compared with pure ZnO [40]. However, it is a well-known fact that structural, electrical, optical, luminescence, and magnetic properties of Co-doped ZnO are strongly dependent on the synthesis, doping, and processing techniques. Co is considered as one of the most effective elements to be doped in ZnO. In this regard, Lu et al. [41] hydrothermally prepare Co-doped ZnO nanorods to study their photocatalytic degradation. Kalpana et al. [42] prepare Co-doped ZnO by co-precipitation method and also study their photocatalytic degradation. Some other synthesis techniques have also been reported in the past, including polymeric sol-gel, polymeric precursor method, co-precipitation, auto combustion method, RF magnetron sputtering, mechanical synthesis, solvothermal synthesis, acrylamide polymerization synthesis, and thermal decomposition, to synthesize ZnO and Co-doped ZnO [43,44,45,46,47]. The impact on the band gap, in particular, its association with near band edge (NBE) emissions in Zn1−xCoxO is still unclear, as many literature studies have provided conflicting results [44]. The environmental exposure of Co is extremely high and location dependent, making it generally difficult to measure. Dietary consumption is thought to be the most common route of exposure for the general public. Cobalt is present in almost all nutrition, with the exception of vitamin B12 and other supplements. Background Co (blood) level is calculated using a normal dietary Co exposure; based on this, it is thought that it will not pose a threat to human health. Many questions about the dose-response characteristics of Co-related adverse health effects have been addressed by a newly developed biokinetic model, which shows that less than 300 µg/L blood Co concentrations are unlikely to cause clinically significant indications in healthy people. Furthermore, regular exposure at acceptable doses is unlikely to cause serious health problems [48,49,50].
Likewise, many researchers have previously investigated the bactericidal properties of ZnO nanostructures in B. subtilis, S. dysenteriae, V. cholerae, E. coli, S. aureus, and S. Typhi [50]. It has been found that the chemical interactions between membrane proteins and nanomaterials, as well as the creation of free radicals due to ZnO-NPs, might be the reason for extraordinary bactericidal properties of undoped and Co-doped ZnO-NPs. In comparison to previous work on Co-doped ZnO nanostructures [51], including nanoparticles [45,52] and nanorods [53], the current work is unique not only for the type of precursors and experimental procedures, but also for the size and cylindrical morphology of the as-obtained Co-doped ZnO, as shown in Table 1, and their effective use as a strong antibacterial agent against two (gram-positive and gram-negative) bacterial strains.

2. Materials and Methods

2.1. Materials

Zinc chloride (ZnCl2), cobalt-II chloride hexahydrate (CoCl2·6H2O), and potassium hydroxide (KOH) were procured from Sigma Aldrich, St. Louis, MO, USA. Reagents (99% analytically graded) were used as received without further refinement. The surface morphology of Co-doped ZnO-MCs was examined with a (QUANTA 250 FEI, FEI Company, Hillsboro, OR, USA) FE-SEM. Structural analysis of Co-doped ZnO were performed using X-ray diffractometer (Ultima IV R.I.C Tokyo, Japan), recorded at 20–80° at CuK radiation (λ = 1.54056 Å). To study the elemental composition, X-ray photoelectron spectroscopy (Thermo specific model K-α) was used. Ultraviolet–visible (UV–Vis, model T-60 Oasis Scientific Inc, Taylors, SC, USA) spectroscopy was used to study optical properties. Four separate bacterial strains were tested for antibacterial activity, namely, gram-negative; Escherichia coli (ATCC® 33876), Klebsiella pneumonia (ATCC® BAA-1144) and gram-positive; Staphylococcus aureus (ATCC® 11632) and Streptococcus pyogenes (ATCC® 19615). Nutrient agar (Oxoid® CM0003) was procured from Sigma-Aldrich (St. Louis, MO, USA).

2.2. Synthesis of Co-Doped ZnO Microcrystals

Highly crystalline ZnO cylindrical microcrystals with Co doping were synthesized hydrothermally by using zinc chloride (ZnCl2), cobalt-II chloride hexahydrate (CoCl2·6H2O), and potassium hydroxide (KOH) as precursors. At first, 1.6 g of ZnCl2 and 0.4 g of CoCl2·6H2O were mixed. Cobalt II chloride hexahydrate is equal to the 25% of the total weight of zinc chloride. This mixture was then dissolved in 40 mL of distilled water (DI). Subsequently, 0.8 g KOH aqueous solution was added dropwise to get a precipitate with a pale pink color. Consequently, at room temperature, the solution was stirred for 30 min then the homogenous mixture was poured in an autoclave (Teflon lined) and kept in an oven for 23 h at 120 °C. The as-obtained precipitate was first washed numerous times with distilled water (DI) and then with ethanol and finally dried at 90 °C, for 30 min. The schematic representation of the whole process is demonstrated in Figure 1.

2.3. Screening of Antibacterial Activity

To assess the efficiency (antibacterial) of Co-doped ZnO-MCs in Mueller–Hinton broth media (from their pure cultures), all bacterial strains were sub-cultured and incubated overnight. Fresh cultures were used in the antibacterial assay by moving stock suspensions mounted on nutrient agar and incubated at 37 °C for a complete day. The bacterial culture turbidity was measured by using the 0.5 McFarland standard [58], which equals 1.5 × 108 (CFU/mL) bacteria. A sterile glass spreader was used to spread each species on a Mueller–Hinton Petri dish. Wells (4 mm) were made by using a sterile polystyrene tip. Co-doped ZnO-MCs were prepared in 2% hydrochloric acid (HCL) with different concentrations (0.1, 0.5, 1 mg/mL). Forty microliters (40 µL) taken from the prepared stock solution was added to each well. Plates were placed entirely for incubation at 37 °C in an incubator overnight. In a UV transilluminator, photoactivation of Co-doped ZnO-MCs suspensions was carried out by exposure at 254 nm with UV light, for 30 min. The inhibition zone was measured around each well in millimeters by a caliper. Clindamycin phosphate, as a reference antibiotic (standard), was used at a concentration of 20 µg/mL. Every experiment was performed three times and mean value was calculated.

3. Results and Discussion

Figure 2a–c shows the field emission scanning electron microscopy (FE-SEM) results for the size and morphology of the as-synthesized Co-doped ZnO-MCs. Figure 2a shows the lower magnification FE-SEM micrograph. Here, individual or isolated microcrystals can be seen or observed randomly aligned in different directions. Along with randomly aligned microcrystals, some smaller species or undeveloped structures are also visible. Figure 2b shows the high magnification FE-SEM micrograph to observe the smaller as well as the larger size microcrystals. The larger size crystals are found to have a length of up to 30 μm, whereas the smaller size crystal can be found in the range of a few to greater than 5 μm. The smaller size crystals are mostly the undeveloped species observed in the previous micrograph in lower magnification. However, some undeveloped species can still be seen. Figure 2c shows the higher magnification micrograph. It confirmed the similar cylindrical morphology of all the crystals in the sample. The larger size cylinder has a diameter of 2 µm, whereas the smaller size is found to have a diameter in the range of 500 nm–1 µm. Smaller size particle-like structures can be seen stuck with the fine and smooth surface of the crystals. It points towards the stages-wise growth and Co-doping of the as-synthesized crystals. The growth and doping start from smaller particle-like structures and developed into smaller cylindrical structures with their condensation due to the increased growth duration. This condensation of the Co-doped ZnO continues with a longer growth duration of 23 h. It means that the size of the crystals can easily be adjusted by optimization of the growth duration.
XRD pattern of Co-doped ZnO-MCs is observed and shown in Figure 3. It is an analytical technique used to determine the size and nature of the material. Intensities and position of the peaks were used to assess the crystallite size, structure, and phase of the material. The miller indices have been calculated for each diffraction peak that confirms the formation of Co-doped ZnO-MCs in the sample. Miller indices for each diffraction peak is (100), (002), (101), (102), (110), (103), (200), (112), (201), (004), and (202), respectively. Crystalline phases were identified using X’Pert HighScore (version: 2.0a (2.0.1), year: 2004, manufacturer: PANalytical B.V., Almelo, Netherlands) and compare the diffraction pattern of the sample with the reference database from inorganic crystal structure database (ICSD) card/ file no: 01-076-0704. Some low intensity peaks were also observed in the sample, which is attributed to the existence of Co-based oxides (Co2O3 and CoO). The antiferromagnetic nature of Co2O3 results in a decrease in Co atomic percentage and sample’s magnetization. Cubic rock salt and hexagonal wurtzite are the two stable phases of cobalt (II) oxide CoO. Given that Zn2+ and Co2+ have ionic radii of 0.74 and 0.745, and that CoO can also crystallize in the hexagonal structure, doping and the change in NP size should have no significant impact on the lattice parameters. A change in parameters is observed due to an increase in Co concentration in ZnO [59].
The synthesized cylindrical microcrystals of Co-doped ZnO were characterized via XPS to confirm its elemental contents and chemical bonding states. The XPS results are displayed in Figure 4a–d. The full range XPS survey for the synthesized microcrystals shows several peaks (tagged for their respective elements: Zn, C, O, and Co) at different values of binding energies. The C 1s peak at 285 eV corresponds to the unavoidable carbon contamination, which might have occurred due to exposure of the sample in the air before its XPS characterization. The high-resolution XPS spectra for Zn 2p peaks are shown in Figure 4b. The first peak observed at 1021.5 eV corresponds to Zn 2p3/2, whereas the second peak observed at 1044.5 eV corresponds to Zn 2p1/2, respectively. Both the peaks in the Zn 2p states have a difference of 23 eV, which, according to the available literature, confirms the Zn2+ chemical states [53,60]. Figure 4c shows the high-resolution O 1 s spectrum with a single peak centered at 530.2 eV. It corresponds to lattice oxygen surrounded by zinc and cobalt ion in the hexagonal wurtzite structure [60]. Similarly, the Co 2p XPS wide scan is shown in Figure 4d with Co 2p3/2 peak centered at 780.5 eV and Co 2p1/2 peak at 796.5 eV. Both Co 2p peaks have a binding energy difference of 16 eV, which corresponds to the existence of Co2+. Thus, the XPS analysis demonstrates the doping of Co2+ in ZnO lattice by the substitution of Zn2+ with no further impurities [53].
Defects, dopant incorporation, and lattice disorder in the as-synthesized Co-doped ZnO-MCs lattice have been analyzed via non-destructive Raman spectroscopy [61]. The as-obtained Raman spectrum shown in Figure 5 has two main peaks at 97.5 and 433.5 cm−1, respectively. The above-mentioned peaks correspond to nonpolar E2 (low) and E2 (high) optical phonon modes of Co-doped ZnO. Unlike pure ZnO, these peaks are shifted towards higher frequencies [62]. The high-intensity peak at 382.5 cm−1 is assigned to the A1 (TO) mode of vibration [63]. The peaks observed at 162 cm−1 can be assigned to defect-induced mode [64], whereas the one seen at 282 cm−1 corresponds to B1 (low) phonons [65,66]. The Raman spectrum of the Co-doped ZnO also has a peak at 538.5 cm−1, which according to the available literature is persuaded by the host lattice defects such as Zn interstitials and oxygen vacancies. This, in other words, easily justifies the Co2+ doping into ZnO lattice [63,67].
Figure 6 shows UV-vis absorption spectra of Co-doped ZnO-MCs. The absorption properties are recorded in the 200–800 nm range. The value of absorption is solely determined by various types of factors such as the size and defects in grain structure. At around 374 nm, there is a rapidly rising absorption edge due to exciton recombination or defects. At higher levels of cobalt doping, the rate of absorption increases. The increase in absorption in the visible region depends on the increase in the concentration of defects causing deep levels in the ZnO band gap [68]. The increase in absorption of light is observed due to increase in lattice defects by cobalt concentration and replacing Zn2+ ions with Co2+ ions in the ZnO lattices [69].
Figure 7 shows PL spectra of Co-doped ZnO-MCs. The PL spectra show a UV near-band edge emission and blue-green emission peaks of around 373 nm and 483 nm wavelengths, respectively. The emission peak (~380 nm) is due to band-to-band excitons’ transition, while the peak at 485 nm is caused by the transition of electrons from the level of the ionized oxygen vacancies to the valence band [70]. Peak observed at ~373 nm is due to the emission from the band edge by radiative annihilation of excitons. These are linked to the recombination of excitons that are both free and shallowly bound [71,72]. The peak observed appearing at 483 nm is attributed to the formation of hydroxyl radicals and surface defects.
The antibacterial efficacy of synthesized Co-doped ZnO-MCs is assessed by agar method [73] against both gram-negative and gram-positive strains. The ability of ZnO materials to kill bacteria is normally determined by reactive oxygen species [74]. The hydrogen peroxide molecules, hydroxyl radical, and superoxide belong to the group of reactive oxygen species (ROS), which not only can cause DNA damage, but also can cause cell death [75,76]. Photocatalytic formation of ROS was the main factor in the antibacterial activity of various metal oxides [77]. Raghupati et al. [74] showed that an increase in the antibacterial activity of ZnO was associated with an increase in the production of ROS from ZnO under the influence of UV radiation. It is known that ZnO has a band gap energy of about 3.2 eV and consequently its excitation is limited to the UV radiation range. Electron-hole pairs are formed when ZnO nanostructures are exposed to UV or visible light. From Co-doped ZnO-MC suspension, these photo-generated holes separated water molecules into H+ and OH- ions. Oxygen species (dissolved) are condensed to superoxide anions (•O-2), which react with H+ to generate (HO2•). The hydrogen peroxide (H2O2) molecules are formed as they collide with the hydrogen ions and majority charge carriers (e-). The hydrogen peroxide molecules produced can penetrate the cytoplasmic membrane, killing the bacteria [78,79,80]. The schematic illustration for the antibacterial mechanism of Co-doped ZnO-MCs is demonstrated in Figure 8. ROS in large amount is produced during various photocatalytic processes. As a result, subcellular damage, including DNA damage, membrane damage, and protein denaturation, appear. Co-doped ZnO-MCs have greater antibacterial activity as a result of improved binding forces and the formation of free radicals in the cell [81,82].
All bacterial strains in the current study were recognized by using different biochemical tests conferring to the described method [83]. Pure bacteria culture was stored in a freeze-dried atmosphere at 4 °C in agar slants until further use. Figure 9 and Figure 10 show the antibacterial effects of the pure and Co-doped ZnO-MCs, against four strains, of which two are gram-negative (E. coli, K. pneumonia) and two are gram-positive (S. aureus, S. pyogenes). The bacterial isolates were treated with varying doses of ZnO and Co-doped ZnO-MCs (0.1, 0.5, and 1 mg/mL) dissolved in HCl (3%). Our outcomes show that, for all the tested doses, the growth of all the microbes is inhibited by both ZnO and Co-doped ZnO-MCs. An increase in ZOI with the increase in the Co-doped ZnO-MCs concentration is observed in Figure 11a,b. According to our results, gram-negative microbes, in comparison with gram-positive microbes, are more sensitive to Co-doped ZnO-MCs. Among gram-negative microbes, E. coli forms 17 ± 0.34 mm and 13 ± 0.26 mm ZOI, whereas K. pneumonia (which is more sensitive to Co-doped ZnO-MCs treatment) forms 19 ± 0.38 mm and 14 ± 0.28 mm ZOI, as shown in Table 2. Among gram-positive microbes, S. pyogenes, forms a ZOI of 16 ± 0.32 mm and 9 ± 0.18 mm, whereas S. aureus forms a ZOI of 15 ± 0.30 mm and 13 ± 0.26 mm on the same dose. Bacterial resistance is found to improve significantly with the increase in the Co content. These results suggest that Co-doped ZnO-MCs can be used as an antibacterial agent in a variety of pharmaceutical and biological applications in the future.

4. Conclusions

A simple and cost-effective hydrothermal technique is used to synthesize the unique cylindrical Co-doped ZnO-MCs. XRD, XPS, and Raman characterizations effectively confirmed the incorporation of Co+2 ions into the Zn+2 ions lattice site. UV-vis results showed that, at a higher dopant concentration (0.4 g/0.025 M), the rate of absorption increases. The increase in absorption in the visible region depends on the increase in the concentration of defects causing deep levels in the ZnO band gap, and this broad absorption is due to d–d transition of Co2+ ions. Co-doped ZnO-MCs can efficiently work against gram-negative and gram-positive bacteria. The results showed that Co-doped ZnO-MCs have better antibacterial activity against gram-negative bacteria, compared with gram-positive bacteria. All the tested bacteria were inhibited by Co-doped ZnO-MCs and the inhibitory effect was dose-dependently increased. Gram-negative microbes were shown to be more sensitive to Co-doped ZnO-MCs, as compared with gram-positive microbes.

Author Contributions

Conceptualization, A.K. (Awais Khalid), P.A., A.I.A., S.M., and M.U.K.; M.R.I.F., M.A.A., A.A.A., and D.A.B.; Data curation, A.K. (Awais Khalid) and P.A.; Formal analysis, A.K. (Awais Khalid), P.A., A.I.A., S.M., A.K. (Abdulhameed Khan), I.U.D., K.A., and D.A.B.; Funding acquisition, M.U.K.; M.R.I.F., K.A., and A.A.A.; Investigation, A.K. (Awais Khalid), P.A., S.M., and M.U.K.; M.R.I.F., A.K. (Abdulhameed Khan), I.U.D., M.A.A., K.A., A.A.A., and D.A.B.; Methodology, A.K. (Awais Khalid), P.A., S.M., A.K. (Abdulhameed Khan), I.U.D., and M.A.A.; Project administration, M.U.K.; M.R.I.F., K.A., A.A.A., and D.A.B.; Resources, P.A., A.I.A., M.U.K., I.U.D., M.A.A., K.A., A.A.A., and D.A.B.; Software, A.I.A., S.M., M.U.K., A.K. (Abdulhameed Khan), K.A., and D.A.B.; Supervision, P.A. and S.M.; Validation, P.A., A.I.A., M.U.K., and D.A.B.; Visualization, I.U.D. and M.A.A.; Writing—original draft, A.K. (Awais Khalid) and P.A.; Writing—review and editing, A.K. (Awais Khalid), P.A., A.I.A., S.M., and M.U.K.; M.R.I.F., A.K. (Abdulhameed Khan), I.U.D., M.A.A., K.A., A.A.A., and D.A.B. All authors have read and agreed to the published version of the manuscript.

Funding

The authors extend their appreciation to the International Scientific Partnership Program ISPP at King Saud University for funding this research through ISPP-20154(1).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the data is available within the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bhushan, B. Fundamentals of Tribology and Bridging the Gap between the Macro-and Micro/Nanoscales; Springer: Berlin/Heidelberg, Germany, 2012. [Google Scholar]
  2. Khalid, P.A.; Ahmad, A.I.; Alharthi, S.; Muhammad, M.U.; Khandaker, M.; Rehman, M.R.I.; Faruque, I.U.; Din, M.A.; Alotaibi, K.A. Structural, optical and antibacterial efficacy of pure and zinc-doped copper oxide against pathogenic bacteria. Nanomaterials 2021, 11, 451. [Google Scholar] [CrossRef]
  3. Meulenkamp, E.A. Synthesis and growth of ZnO nanoparticles. J. Phys. Chem. B 1998, 102, 5566–5572. [Google Scholar] [CrossRef]
  4. Özgür, Ü.; Alivov, Y.I.; Teke, C.L.A.; Reshchikov, M.A.; Doğan, S.; Avrutin, V.; Cho, S.-J.; Morkoç, H. A comprehensive review of ZnO materials and devices. J. Appl. Phys. 2005, 98, 11. [Google Scholar] [CrossRef] [Green Version]
  5. Ahmad, P.; Khandaker, M.U.; Amin, Y.M.; Muhammad, N. Synthesis and characterization of boron nitride microtubes. Mater. Express 2015, 5, 249–254. [Google Scholar] [CrossRef]
  6. Ozgür, Ü.; Hofstetter, D.; Morkoç, H. ZnO devices and applications: A review of current status and future prospects. Proc. IEEE 2010, 98, 1255–1268. [Google Scholar] [CrossRef]
  7. Necib, K.; Touam, T.; Chelouche, A.; Ouarez, L.; Djouadi, D.; Boudine, B. Investigation of the effects of thickness on physical properties of AZO sol-gel films for photonic device applications. J. Alloys Compd. 2018, 735, 2236–2246. [Google Scholar] [CrossRef]
  8. Srivastava, A.; Kumar, N.; Misra, K.P.; Khare, S. Enhancement of band gap of ZnO nanocrystalline films at a faster rate using Sr dopant. Electron. Mater. Lett. 2014, 10, 703–711. [Google Scholar] [CrossRef]
  9. Atif, M.; Younas, U.; Khalid, W.; Ahmed, Z.; Ali, M.; Nadeem, Z. Impedance spectroscopy, ferroelectric and optical properties of cobalt-doped Zn1−xCoxO nanoparticles. J. Mater. Sci. 2020, 31, 1–9. [Google Scholar]
  10. Chattopadhyay, S.; Misra, K.P.; Agarwala, A.; Rao, A.; Babu, P. Correlated quartic variation of band gap and NBE energy in sol-gel derived Zn1−xCoxO nanoparticles. Mater. Chem. Phys. 2019, 227, 236–241. [Google Scholar] [CrossRef]
  11. Nair, M.G.; Nirmala, M.; Rekha, K.; Anukaliani, A. Structural, optical, photo catalytic and antibacterial activity of ZnO and Co doped ZnO nanoparticles. Mater. Lett. 2011, 65, 1797–1800. [Google Scholar] [CrossRef]
  12. Chithira, P.; John, T.T. Correlation among oxygen vacancy and doping concentration in controlling the properties of cobalt doped ZnO nanoparticles. J. Magn. Magn. Mater. 2020, 496, 165928. [Google Scholar] [CrossRef]
  13. Thool, G.S.; Singh, A.K.; Singh, R.; Gupta, A.; Susan, A.B.H. Facile synthesis of flat crystal ZnO thin films by solution growth method: A micro-structural investigation. J. Saudi Chem. Soc. 2014, 18, 712–721. [Google Scholar] [CrossRef] [Green Version]
  14. Shatnawi, M.; Alsmadi, A.; Bsoul, I.; Salameh, B.; Alna’Washi, G.; Al-Dweri, F.; El Akkad, F. Magnetic and optical properties of Co-doped ZnO nanocrystalline particles. J. Alloys Compd. 2016, 655, 244–252. [Google Scholar] [CrossRef]
  15. Sharma, D.; Jha, R. Transition metal (Co, Mn) co-doped ZnO nanoparticles: Effect on structural and optical properties. J. Alloys Compd. 2017, 698, 532–538. [Google Scholar] [CrossRef]
  16. Goel, S.; Kumar, B. A review on piezo-/ferro-electric properties of morphologically diverse ZnO nanostructures. J. Alloys Compd. 2020, 816, 152491. [Google Scholar] [CrossRef]
  17. Khalid, A.; Ahmad, P.; Alharthi, A.; Muhammad, S.; Khandaker, M.; Faruque, M.I.; Din, I.; Alotaibi, M. Unmodified titanium dioxide nanoparticles as a potential contrast agent in photon emission computed tomography. Crystals 2021, 11, 171. [Google Scholar] [CrossRef]
  18. Fukumura, T.; Jin, Z.; Ohtomo, A.; Koinuma, H.; Kawasaki, M. An oxide-diluted magnetic semiconductor: Mn-doped ZnO. Appl. Phys. Lett. 1999, 75, 3366–3368. [Google Scholar] [CrossRef]
  19. Dejene, F.; Onani, M.; Koao, L.; Wako, A.; Motloung, S.; Yihunie, M. Structure, morphology and optical properties of undoped and MN-doped ZnO(1−x)Sx nano-powders prepared by precipitation method. Physica B 2016, 480, 63–67. [Google Scholar] [CrossRef]
  20. Godavarthi, U.; Mote, V.; Reddy, M.R.; Nagaraju, P.; Kumar, Y.V.; Dasari, K.T.; Dasari, M.P. Precipitated cobalt doped ZnO nanoparticles with enhanced low temperature xylene sensing properties. Physica B 2019, 553, 151–160. [Google Scholar] [CrossRef]
  21. Arif, M.; Sanger, A.; Shkir, M.; Singh, A.; Katiyar, R. Influence of interparticle interaction on the structural, optical and magnetic properties of NiO nanoparticles. Physica B 2019, 552, 88–95. [Google Scholar] [CrossRef]
  22. Khalid, A.; Ahmad, P.; Alharthi, A.I.; Muhammad, S.; Khandaker, M.U.; Faruque, M.R.I.; Din, I.U.; Alotaibi, M.A.; Khan, A. Synergistic effects of Cu-doped ZnO nanoantibiotic against Gram-positive bacterial strains. PLoS ONE 2021, 16, e0251082. [Google Scholar] [CrossRef] [PubMed]
  23. Ahmad, P.; Khandaker, M.U.; Muhammad, N.; Khan, G.; Rehman, F.; Khan, A.S.; Ullah, Z.; Khan, A.; Ali, H.; Ahmed, S.M.; et al. Fabrication of hexagonal boron nitride quantum dots via a facile bottom-up technique. Ceram. Int. 2019, 45, 22765–22768. [Google Scholar] [CrossRef]
  24. Ahmed, S.M.; Kazi, S.N.; Khan, G.; Dahari, M.; Zubir, M.N.M.; Ahmad, P.; Montazer, E. Experimental investigation on momentum and drag reduction of Malaysian crop suspensions in closed conduit flow. IOP Conf. Ser. 2017, 210, 012065. [Google Scholar] [CrossRef]
  25. Ahmad, P.; Khandaker, M.U.; Amin, Y.M.; Muhammad, N.; Khan, G.; Khan, A.S.; Numan, A.; Rehman, M.A.; Ahmed, S.M.; Khan, A. Synthesis of hexagonal boron nitride fibers within two hour annealing at 500 °C and two hour growth duration at 1000 °C. Ceram. Int. 2016, 42, 14661–14666. [Google Scholar] [CrossRef]
  26. Shannon, R.D. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallogr. 1976, 32, 751–767. [Google Scholar] [CrossRef]
  27. Hammad, T.M.; Salem, J.K.; Harrison, R.G. Structure, optical properties and synthesis of Co-doped ZnO superstructures. Appl. Nanosci. 2013, 3, 133–139. [Google Scholar] [CrossRef] [Green Version]
  28. Straumal, B.; Baretzky, B.; Mazilkin, A.; Protasova, S.; Myatiev, A.; Straumal, P. Increase of Mn solubility with decreasing grain size in ZnO. J. Eur. Ceram. Soc. 2009, 29, 1963–1970. [Google Scholar] [CrossRef]
  29. Woo, H.-S.; Kwak, C.-H.; Chung, J.-H.; Lee, J.-H. Co-doped branched ZnO nanowires for ultraselective and sensitive detection of xylene. ACS Appl. Mater. Interfaces 2014, 6, 22553–22560. [Google Scholar] [CrossRef]
  30. Zhu, S.; Xu, L.; Yang, S.; Zhou, X.; Chen, X.; Dong, B.; Bai, X.; Lu, G.; Song, H. Cobalt-doped ZnO nanoparticles derived from zeolite imidazole frameworks: Synthesis, characterization, and application for the detection of an exhaled diabetes biomarker. J. Colloid Interface Sci. 2020, 569, 358–365. [Google Scholar] [CrossRef]
  31. Taskin, M.; Podder, J. Structural, optical and electrical properties of pure and Co-doped ZnO nano fiber thin films prepared by spray pyrolysis. Appl. Sci. Rep. 2014, 2, 107–113. [Google Scholar]
  32. Adeel, M.; Saeed, M.; Khan, I.; Muneer, M.; Akram, N. Synthesis and characterization of Co-ZnO and evaluation of its photocatalytic activity for photodegradation of methyl orange. ACS Omega 2021, 6, 1426–1435. [Google Scholar] [CrossRef] [PubMed]
  33. Fabbiyola, S.; Kennedy, L.J. Bandgap engineering in doped ZnO nanostructures for dye sensitized solar cell applications. J. Nanosci. Nanotechnol. 2019, 19, 2963–2970. [Google Scholar] [CrossRef] [PubMed]
  34. Ali, H.T.; Jacob, J.; Khalid, M.; Mahmood, K.; Yusuf, M.; Mehboob, K.; Ikram, S.; Ali, A.; Amin, N.; Ashar, A. Optimizing the structural, morphological and thermoelectric properties of zinc oxide by the modulation of cobalt doping concentration. J. Alloys Compd. 2021, 871, 159564. [Google Scholar] [CrossRef]
  35. Kumar, R.; Kumar, P. Co-Doped ZnO Nanoparticles Reinforcement in PVDF for 3D Printing of Magnetic Structures; Elsevier: Amsterdam, The Netherlands, 2020. [Google Scholar] [CrossRef]
  36. Skoda, D.; Urbanek, P.; Sevcik, J.; Münster, L.; Nadazdy, V.; Cullen, D.A.; Bazant, P.; Antos, J.; Kuritka, I. Colloidal cobalt-doped ZnO nanoparticles by microwave-assisted synthesis and their utilization in thin composite layers with MEH-PPV as an electroluminescent material for polymer light emitting diodes. Org. Electron. 2018, 59, 337–348. [Google Scholar] [CrossRef]
  37. Zang, W.; Li, P.; Fu, Y.; Xing, L.; Xue, X. Hydrothermal synthesis of Co-ZnO nanowire array and its application as piezo-driven self-powered humidity sensor with high sensitivity and repeatability. RSC Adv. 2015, 5, 84343–84349. [Google Scholar] [CrossRef]
  38. Vijayaprasath, G.; Murugan, R.; Narayanan, J.S.; Dharuman, V.; Ravi, G.; Hayakawa, Y. Glucose sensing behavior of cobalt doped ZnO nanoparticles synthesized by co-precipitation method. J. Mater. Sci. Mater. Electron. 2015, 26, 4988–4996. [Google Scholar] [CrossRef]
  39. Kołodziejczak-Radzimska, A.; Jesionowski, T. Zinc oxide—From synthesis to application: A review. Materials 2014, 7, 2833–2881. [Google Scholar] [CrossRef] [Green Version]
  40. Kumaresan, S.; Vallalperuman, K.; Sathishkumar, S.; Karthik, M.; Siva-Karthik, P. Synthesis and systematic investigations of Al and Cu-doped ZnO nanoparticles and its structural, optical and photo-catalytic properties. J. Mater. Sci. 2017, 28, 9199–9205. [Google Scholar] [CrossRef]
  41. Lu, Y.; Lin, Y.; Wang, D.; Wang, L.; Xie, T.; Jiang, T. A high performance cobalt-doped ZnO visible light photocatalyst and its photogenerated charge transfer properties. Nano Res. 2011, 4, 1144–1152. [Google Scholar] [CrossRef]
  42. Kalpana, S.; Krishnan, S.; Senthil, T.; Elangovan, S. Cobalt doped Zinc oxide nanoparticles for photocatalytic applications. J. Ovonic Res. 2017, 13, 263–269. [Google Scholar]
  43. Wang, X.L.; Luan, C.Y.; Shao, Q.; Pruna, A.; Leung, C.W.; Lortz, R.; Zapien, J.A.; Ruotolo, A. Effect of the magnetic order on the room-temperature band-gap of Mn-doped ZnO thin films. Appl. Phys. Lett. 2013, 102, 102112. [Google Scholar] [CrossRef] [Green Version]
  44. Srinet, G.; Varshney, P.; Kumar, R.; Sajal, V.; Kulriya, P.; Knobel, M.; Sharma, S. Structural, optical and magnetic properties of Zn1-xCoxO prepared by the sol-gel route. Ceram. Int. 2013, 39, 6077–6085. [Google Scholar] [CrossRef]
  45. Gandhi, V.; Ganesan, R.; Syedahamed, H.H.A.; Thaiyan, M. Effect of cobalt doping on structural, optical, and magnetic properties of ZnO nanoparticles synthesized by coprecipitation method. J. Phys. Chem. C 2014, 118, 9715–9725. [Google Scholar] [CrossRef]
  46. Djerdj, I.; Jagličić, Z.; Arčon, D.; Niederberger, M. Co-doped ZnO nanoparticles: Minireview. Nanoscale 2010, 2, 1096–1104. [Google Scholar] [CrossRef]
  47. Wojnarowicz, J.; Chudoba, T.; Lojkowski, W. A review of microwave synthesis of zinc oxide nanomaterials: Reactants, process parameters and morphologies. Nanomaterials 2020, 10, 1086. [Google Scholar] [CrossRef] [PubMed]
  48. Leyssens, L.; Vinck, B.; van der Straeten, C.; Wuyts, F.; Maes, L. Cobalt toxicity in humans—A review of the potential sources and systemic health effects. Toxicology 2017, 387, 43–56. [Google Scholar] [CrossRef]
  49. Tvermoes, B.E.; Paustenbach, D.J.; Kerger, B.D.; Finley, B.L.; Unice, K.M. Review of cobalt toxicokinetics following oral dosing: Implications for health risk assessments and metal-on-metal hip implant patients. Crit. Rev. Toxicol. 2015, 45, 367–387. [Google Scholar] [CrossRef]
  50. Unice, K.M.; Kerger, B.D.; Paustenbach, D.J.; Finley, B.L.; Tvermoes, B.E. Refined biokinetic model for humans exposed to cobalt dietary supplements and other sources of systemic cobalt exposure. Chem. Biol. Interact. 2014, 216, 53–74. [Google Scholar] [CrossRef]
  51. Kaphle, A.; Reed, T.; Apblett, A.; Hari, P. Doping efficiency in cobalt-doped ZnO nanostructured materials. J. Nanomater. 2019, 2019, 1–13. [Google Scholar] [CrossRef]
  52. Rajendar, V.; Dayakar, T.; Shobhan, K.; Srikanth, I.; Rao, K.V. Systematic approach on the fabrication of Co doped ZnO semiconducting nanoparticles by mixture of fuel approach for Antibacterial applications. Superlattices Microstruct. 2014, 75, 551–563. [Google Scholar] [CrossRef]
  53. Narasimman, S.; Balakrishnan, L.; Alex, Z.C. Fiber optic magnetic field sensor using Co doped ZnO nanorods as cladding. RSC Adv. 2018, 8, 18243–18251. [Google Scholar] [CrossRef] [Green Version]
  54. Rana, S.B.; Singh, R.P.P.; Arya, S. Structural, optical, magnetic and antibacterial study of pure and cobalt doped ZnO nanoparticles. J. Mater. Sci. Mater. Electron. 2016, 28, 2660–2672. [Google Scholar] [CrossRef]
  55. Godavarthi, U.; Mote, V.; Dasari, M. Role of cobalt doping on the electrical conductivity of ZnO nanoparticles. J. Asian Ceram. Soc. 2017, 5, 391–396. [Google Scholar] [CrossRef]
  56. Kayani, Z.N.; Shah, I.; Zulfiqar, B.; Riaz, S.; Naseem, S.; Sabah, A. Structural, optical and magnetic properties of nanocrystalline co-doped ZnO thin films grown by sol-gel. Zeitschrift für Naturforschung A 2017, 73, 13–21. [Google Scholar] [CrossRef]
  57. Kaphle, A.; Echeverria, E.; Mcllroy, D.N.; Roberts, K.; Hari, P. Thermo-optical properties of cobalt-doped zinc oxide (ZnO) nanorods. J. Nanosci. Nanotechnol. 2019, 19, 3893–3904. [Google Scholar] [CrossRef]
  58. National Committee for Clinical Laboratory Standards; Barry, A.L. Methods for Determining Bactericidal Activity of Antimicrobial Agents: Approved Guideline; Clinical and Laboratory Standards Institute: Wayne, PA, USA, 1999. [Google Scholar]
  59. Wojnarowicz, J.; Kusnieruk, S.; Chudoba, T.; Gierlotka, S.; Lojkowski, W.; Knoff, W.; Lukasiewicz, M.I.; Witkowski, B.S.; Wolska, A.; Klepka, M.T.; et al. Paramagnetism of cobalt-doped ZnO nanoparticles obtained by microwave solvothermal synthesis. Beilstein J. Nanotechnol. 2015, 6, 1957–1969. [Google Scholar] [CrossRef] [Green Version]
  60. Hu, J.; Gao, F.; Zhao, Z.; Sang, S.; Li, P.; Zhang, W.; Zhou, X.; Chen, Y. Synthesis and characterization of Cobalt-doped ZnO microstructures for methane gas sensing. Appl. Surf. Sci. 2016, 363, 181–188. [Google Scholar] [CrossRef]
  61. Singhal, A.; Achary, S.N.; Manjanna, J.; Chatterjee, S.; Ayyub, P.; Tyagi, A.K. Chemical synthesis and structural and magnetic properties of dispersible cobalt- and nickel-doped ZnO nanocrystals. J. Phys. Chem. C 2010, 114, 3422–3430. [Google Scholar] [CrossRef]
  62. Taher, F.A.; Abdeltwab, E. Shape-controlled synthesis of nanostructured Co-doped ZnO thin films and their magnetic properties. CrystEngComm 2018, 20, 5844–5856. [Google Scholar] [CrossRef]
  63. Pal, B.; Giri, P.K. High temperature ferromagnetism and optical properties of Co doped ZnO nanoparticles. J. Appl. Phys. 2010, 108, 084322. [Google Scholar] [CrossRef] [Green Version]
  64. Bundesmann, C.; Ashkenov, N.; Schubert, M.; Spemann, D.; Butz, T.; Kaidashev, E.M.; Lorenz, M.; Grundmann, M. Raman scattering in ZnO thin films doped with Fe, Sb, Al, Ga, and Li. Appl. Phys. Lett. 2003, 83, 1974–1976. [Google Scholar] [CrossRef]
  65. Manjón, F.J.; Marí, B.; Serrano, J.; Romero, A.H. Silent Raman modes in zinc oxide and related nitrides. J. Appl. Phys. 2005, 97, 053516. [Google Scholar] [CrossRef] [Green Version]
  66. Chanda, A.; Gupta, S.; Vasundhara, M.; Joshi, S.R.; Mutta, G.R.; Singh, J. Study of structural, optical and magnetic properties of cobalt doped ZnO nanorods. RSC Adv. 2017, 7, 50527–50536. [Google Scholar] [CrossRef] [Green Version]
  67. Duan, L.B.; Rao, G.H.; Wang, Y.C.; Yu, J.; Wang, T. Magnetization and Raman scattering studies of (Co, Mn) codoped ZnO nanoparticles. J. Appl. Phys. 2008, 104, 013909. [Google Scholar] [CrossRef]
  68. Ghosh, J.; Ghosh, R.; Giri, P. Tuning the visible photoluminescence in Al doped ZnO thin film and its application in label-free glucose detection. Sens. Actuators B 2018, 254, 681–689. [Google Scholar] [CrossRef]
  69. Çakıcı, T.; Sarıtaş, S.; Muğlu, G.M.; Kundakçı, M.; Yildirim, M. Magnetic, optical and structural characterization of ZnO: Co; ZnO: Fe thin films. AIP Conf. Proc. 2017, 1833, 020094. [Google Scholar]
  70. Lim, J.; Shin, K.; Kim, H.W.; Lee, C. Photoluminescence studies of ZnO thin films grown by atomic layer epitaxy. J. Lumin. 2004, 109, 181–185. [Google Scholar] [CrossRef]
  71. Azmi, R.; Oh, S.-H.; Jang, S.-Y. High-efficiency colloidal quantum dot photovoltaic devices using chemically modified heterojunctions. ACS Energy Lett. 2016, 1, 100–106. [Google Scholar] [CrossRef]
  72. Baiqi, W.; Xudong, S.; Qiang, F.; Iqbal, J.; Yan, L.; Honggang, F.; Dapeng, Y. Photoluminescence properties of Co-doped ZnO nanorods array fabricated by the solution method. Phys. E 2009, 41, 413–417. [Google Scholar] [CrossRef]
  73. Valgas, C.; Souza, S.M.D.; Smânia, E.F.; Smânia, A., Jr. Screening methods to determine antibacterial activity of natural products. Braz. J. Microbiol. 2007, 38, 369–380. [Google Scholar] [CrossRef] [Green Version]
  74. Raghupathi, K.R.; Koodali, R.T. Manna, size-dependent bacterial growth inhibition and mechanism of antibacterial activity of zinc oxide nanoparticles. Langmuir 2011, 27, 4020–4028. [Google Scholar] [CrossRef]
  75. Alfadda, A.A.; Sallam, R.M. Reactive oxygen species in health and disease. J. Biomed. Biotechnol. 2012, 2012, 1–14. [Google Scholar] [CrossRef]
  76. Lee, S.H.; Gupta, M.; Bang, J.B.; Bae, H.; Sung, H.-J. Current progress in reactive oxygen species (ROS)-responsive materials for biomedical applications. Adv. Heal. Mater. 2013, 2, 908–915. [Google Scholar] [CrossRef] [Green Version]
  77. Khalid, A.; Ahmad, P.; Alharthi, A.I.; Mohammad, S.; Khandaker, M.U.; Faruque, M.R.I.; Din, U.; Alotaibi, M.A. A practical method for incorporation of Fe (III) in Titania matrix for photocatalytic applications. Mater. Res. Express 2021. [Google Scholar] [CrossRef]
  78. Semeraro, P.; Bettini, S.; Sawalha, S.; Pal, S.; Licciulli, A.; Marzo, F.; Lovergine, N.; Valli, L.; Giancane, G. Photocatalytic degradation of tetracycline by ZnO/γ-Fe2O3 paramagnetic nanocomposite material. Nanomaterials 2020, 10, 1458. [Google Scholar] [CrossRef]
  79. Bettini, S.; Pagano, R.; Semeraro, P.; Ottolini, M.; Salvatore, L.; Marzo, F.; Lovergine, N.; Giancane, G.; Valli, L. SiO 2—Coated ZnO nanoflakes decorated with Ag nanoparticles for photocatalytic water oxidation. Chem. A Eur. J. 2019, 25, 14123. [Google Scholar] [CrossRef]
  80. Fang, M.; Chen, J.-H.; Xu, X.-L.; Yang, P.-H.; Hildebrand, H.F. Antibacterial activities of inorganic agents on six bacteria associated with oral infections by two susceptibility tests. Int. J. Antimicrob. Agents 2006, 27, 513–517. [Google Scholar] [CrossRef]
  81. Hameed, A.S.H.; Karthikeyan, C.; Sasikumar, S.; Kumar, V.S.; Kumaresan, S.; Ravi, G. Impact of alkaline metal ions Mg2+, Ca2+, Sr2+ and Ba2+ on the structural, optical, thermal and antibacterial properties of ZnO nanoparticles prepared by the co-precipitation method. J. Mater. Chem. B 2013, 1, 5950–5962. [Google Scholar] [CrossRef] [PubMed]
  82. Zhang, L.; Jiang, Y.; Ding, Y.; Daskalakis, N.; Jeuken, L.; Povey, M.; O’Neill, A.J.; York, D.W. Mechanistic investigation into antibacterial behaviour of suspensions of ZnO nanoparticles against E. coli. J. Nanopart. Res. 2010, 12, 1625–1636. [Google Scholar] [CrossRef]
  83. Cowan, S.T. Cowan and Steel’s Manual for the Identification of Medical Bacteria; Cambridge University Press: Cambridge, UK, 2004. [Google Scholar]
Figure 1. Schematic representation for the experimental setup.
Figure 1. Schematic representation for the experimental setup.
Materials 14 03223 g001
Figure 2. (a) Lower, (b) high, and (c) higher magnification micrographs of the as-synthesized Co-doped ZnO cylindrical microcrystals.
Figure 2. (a) Lower, (b) high, and (c) higher magnification micrographs of the as-synthesized Co-doped ZnO cylindrical microcrystals.
Materials 14 03223 g002
Figure 3. XRD pattern of the as-synthesized Co-doped ZnO microcrystals.
Figure 3. XRD pattern of the as-synthesized Co-doped ZnO microcrystals.
Materials 14 03223 g003
Figure 4. (a) XPS survey of Co-doped ZnO microcrystals. (b) High resolution XPS spectra of Zn 2p. (c) High resolution XPS spectrum of O 1s. (d) High resolution XPS spectra of Co 2p.
Figure 4. (a) XPS survey of Co-doped ZnO microcrystals. (b) High resolution XPS spectra of Zn 2p. (c) High resolution XPS spectrum of O 1s. (d) High resolution XPS spectra of Co 2p.
Materials 14 03223 g004
Figure 5. Raman spectrum of Co-doped ZnO microcrystals (MCs).
Figure 5. Raman spectrum of Co-doped ZnO microcrystals (MCs).
Materials 14 03223 g005
Figure 6. UV-vis absorption spectra of cobalt-doped ZnO microcrystals (MCs).
Figure 6. UV-vis absorption spectra of cobalt-doped ZnO microcrystals (MCs).
Materials 14 03223 g006
Figure 7. Photoluminescence spectra of Co-doped ZnO-MCs.
Figure 7. Photoluminescence spectra of Co-doped ZnO-MCs.
Materials 14 03223 g007
Figure 8. Schematic illustration for antimicrobial mechanism of Co-doped ZnO microcrystals against microbial strains.
Figure 8. Schematic illustration for antimicrobial mechanism of Co-doped ZnO microcrystals against microbial strains.
Materials 14 03223 g008
Figure 9. Zone of inhibition (ZOI) formed by ZnO against different bacteria.
Figure 9. Zone of inhibition (ZOI) formed by ZnO against different bacteria.
Materials 14 03223 g009
Figure 10. Zone of inhibition (ZOI) formed by Co-doped ZnO microcrystals against different bacteria.
Figure 10. Zone of inhibition (ZOI) formed by Co-doped ZnO microcrystals against different bacteria.
Materials 14 03223 g010aMaterials 14 03223 g010b
Figure 11. Bar graph displaying the diameter of the ZOI (in mm) produced by (a) ZnO and (b) Co-doped ZnO microcrystals (MCs) against gram-negative and gram-positive bacteria.
Figure 11. Bar graph displaying the diameter of the ZOI (in mm) produced by (a) ZnO and (b) Co-doped ZnO microcrystals (MCs) against gram-negative and gram-positive bacteria.
Materials 14 03223 g011
Table 1. Comparison of the experimental procedure and other parameters of Co-doped ZnO-MCs with the literature.
Table 1. Comparison of the experimental procedure and other parameters of Co-doped ZnO-MCs with the literature.
S. No/ReferencePrecursorsTemperature/TimeTechniqueMorphologyProductConfirmationYear
[52]Zn(NO3)2 6H2O,
Co(NO3)2 6H2O
100 °C/16 hSol-gel combustionGranular surfaceCo-doped-ZnOFE-SEM *2014
[14]ZnO, CoO250 rpm/12 hBall millingNano-particlesCo-doped-ZnOSEM **2016
[54]Zn(CH3COO)2·2H2O, Co(CH3COO)2·4H2ORoom temperature /3 hWet precipitationNano-particlesCo-doped-ZnOSEM2017
[55]Zn(CH3COO)2·2H2O, Co(CH3COO)2·4H2O325 K/2 hCo-PrecipitationNano-particlesCo-doped-ZnOSEM2017
[56]Zn(CH3COO)2·H2O, Co(NO3)2·6H2O60 °C/0.5 hSol-gel dip-coatingClustered grainsCo-doped-ZnOSEM2017
[36]Zn(OAc)2·2H2O,
Co(II)(Acac)2
250 °C/0.25 hMicrowave-assisted polyolNano colloidsCo-doped-ZnOSEM2018
[57]Zn(NO3)2·6H2O,
Co(NO3)3·6H2O
95 °C/6 hChemical bath depositionNano rodsCo-doped-ZnOSEM2019
Our ArticleZnCl2, CoCl3·6H2O120 °C/23 hHydro-thermalCylindrical microcrystalsCo-doped-ZnOFE-SEM2021
* Field emission scanning electron microscope, ** Scanning electron microscope.
Table 2. Summary of the information of bacteria and other findings.
Table 2. Summary of the information of bacteria and other findings.
BacteriaZnOCo Doped ZnO
100
µg/mL
500
µg/mL
1
mg/mL
100
µg/mL
500
µg/mL
1
mg/mL
Gram-negativeE. coliInhibition zone (mm)9 ± 0.1811 ± 0.2613 ± 0.2610 ± 0.213 ± 0.2617 ± 0.34
K. pneumoniae10 ± 0.210 ± 0.2114 ± 0.2811 ± 0.2215 ± 0.319 ± 0.38
Gram-positiveS. aureus9 ± 0.1912 ± 0.2413 ± 0.269 ± 0.1811 ± 0.2215 ± 0.3
S. pyogenes8 ± 0.169 ± 0.189 ± 0.1810 ± 0.213 ± 0.2616 ± 0.32
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Khalid, A.; Ahmad, P.; Alharthi, A.I.; Muhammad, S.; Khandaker, M.U.; Faruque, M.R.I.; Khan, A.; Din, I.U.; Alotaibi, M.A.; Alzimami, K.; et al. Enhanced Optical and Antibacterial Activity of Hydrothermally Synthesized Cobalt-Doped Zinc Oxide Cylindrical Microcrystals. Materials 2021, 14, 3223. https://0-doi-org.brum.beds.ac.uk/10.3390/ma14123223

AMA Style

Khalid A, Ahmad P, Alharthi AI, Muhammad S, Khandaker MU, Faruque MRI, Khan A, Din IU, Alotaibi MA, Alzimami K, et al. Enhanced Optical and Antibacterial Activity of Hydrothermally Synthesized Cobalt-Doped Zinc Oxide Cylindrical Microcrystals. Materials. 2021; 14(12):3223. https://0-doi-org.brum.beds.ac.uk/10.3390/ma14123223

Chicago/Turabian Style

Khalid, Awais, Pervaiz Ahmad, Abdulrahman I. Alharthi, Saleh Muhammad, Mayeen Uddin Khandaker, Mohammad Rashed Iqbal Faruque, Abdulhameed Khan, Israf Ud Din, Mshari A. Alotaibi, Khalid Alzimami, and et al. 2021. "Enhanced Optical and Antibacterial Activity of Hydrothermally Synthesized Cobalt-Doped Zinc Oxide Cylindrical Microcrystals" Materials 14, no. 12: 3223. https://0-doi-org.brum.beds.ac.uk/10.3390/ma14123223

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop