Next Article in Journal
Impurity-Induced Spin-State Crossover in La0.8Sr0.2Co1−xAlxO3
Next Article in Special Issue
Size Control of Ti4O7 Nanoparticles by Carbothermal Reduction Using a Multimode Microwave Furnace
Previous Article in Journal
Microstructure and Mechanical Properties Investigation of the CoCrFeNiNbx High Entropy Alloy Coatings
Previous Article in Special Issue
Current Trends in the Development of Microwave Reactors for the Synthesis of Nanomaterials in Laboratories and Industries: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural and Magnetic Properties of Co‒Mn Codoped ZnO Nanoparticles Obtained by Microwave Solvothermal Synthesis

by
Jacek Wojnarowicz
1,*,
Myroslava Omelchenko
2,
Jacek Szczytko
2,
Tadeusz Chudoba
1,
Stanisław Gierlotka
1,
Andrzej Majhofer
2,
Andrzej Twardowski
2 and
Witold Lojkowski
1
1
Institute of High Pressure Physics, Polish Academy of Sciences, Sokolowska 29/37, 01-142 Warsaw, Poland
2
Institute of Experimental Physics, Faculty of Physics, University of Warsaw, Pasteura 5, 02-093 Warsaw, Poland
*
Author to whom correspondence should be addressed.
Crystals 2018, 8(11), 410; https://doi.org/10.3390/cryst8110410
Submission received: 11 October 2018 / Revised: 27 October 2018 / Accepted: 29 October 2018 / Published: 31 October 2018
(This article belongs to the Special Issue Microwave-Assisted Synthesis of Nanocrystals and Nanostructures)

Abstract

:
Zinc oxide nanoparticles codoped with Co2+ and Mn2+ ions (Zn(1−x−y)MnxCoyO NPs) were obtained for the first time by microwave solvothermal synthesis. The nominal content of Co2+ and Mn2+ in Zn(1−x−y)MnxCoyO NPs was x = y = 0, 1, 5, 10 and 15 mol % (the amount of both ions was equal). The precursors were obtained by dissolving zinc acetate dihydrate, manganese (II) acetate tetrahydrate and cobalt (II) acetate tetrahydrate in ethylene glycol. The morphology, phase purity, lattice parameters, dopants content, skeleton density, specific surface area, average particle size, average crystallite size, crystallite size distribution and magnetic properties of NPs were determined. The real content of dopants was up to 25.0% for Mn2+ and 80.5% for Co2+ of the nominal content. The colour of the samples changed from white to dark olive green in line with the increasing doping level. Uniform spherical NPs with wurtzite structure were obtained. The average size of NPs decreased from 29 nm to 21 nm in line with the increase in the dopant content. Brillouin type paramagnetism and an antiferromagnetic interaction between the magnetic ions was found for all samples, except for that with 15 mol % doping level, where a small ferromagnetic contribution was found. A review of the preparation methods of Co2+ and Mn2+ codoped ZnO is presented.

Graphical Abstract

1. Introduction

Zinc oxide (ZnO) is a multi-functional material [1,2]. The ZnO structure that is thermodynamically stable at room temperature is wurtzite [3]. The ZnO is characterised by a wide band gap of 3.37 eV and a high exciton binding energy (60 meV), which permits the application of ZnO in electronics and optoelectronics [4,5,6,7,8,9,10,11]. The antibacterial and antifungal action of ZnO put ZnO NPs in the centre of interest in the fields of biomedicine and dentistry [11,12,13,14,15,16,17,18,19,20,21,22]. Research is in progress to find new applications of ZnO NPs in producing, for instance, solar cells [23,24], storage media [25], liquid crystals [26], nanostructured polymer composites [27], water filtration [28], photocatalysts [29,30,31], sensing applications [32,33], coatings for UV protection [34,35,36] and crop protection [37,38]. The development of the technologies of synthesis and modification of ZnO properties has made ZnO NPs the focus of spintronics [6,39].
Spintronics is the science of electron spin and the related magnetic moment in semiconductors. Many researchers share the view that the 21st century will be remembered as the age of spintronic revolution due to the construction of the first spintronic devices for quantum computation or communication [40]. Spintronics enables the creation of next-generation optoelectronic devices and data carriers [41]. New materials, e.g., diluted magnetic semiconductors (DMS), where spin-polarised charge carriers can be obtained to enable performance of operations in spintronic devices, are being intensively sought. DMS are materials that are characterised by semiconductor (tunable conductivity) and ferromagnetic (controlled spin polarisation) properties.
Dietl et al. [42] discussed the Zener model of ferromagnetism in zinc-blend magnetic semiconductors including ZnO by calculating Curie temperature Tc higher than 100 K. Doped ZnO has long been considered a promising material for applications as DMS as a result of theoretical calculations, implying that doped ZnO might display ferromagnetic properties at the room temperature [43,44]. Additionally, such ZnO properties as the band gap value and the conductivity could be controlled through doping ZnO NPs with ions of transition metals (e.g., Co, Mn, Cr, Ni, Fe, V). The issue of doped ZnO NPs has been quite extensively examined by various research groups [6,44,45,46,47].
ZnO NPs doped with Co2+ (Zn(1−y)CoyO) and Mn2+ (Zn(1−x)MnxO) ions have gained much interest from scientists [4,6]. Despite numerous studies, the magnetic properties of Zn(1−x)MnxO NPs and Zn(1−y)CoyO NPs have been a fairly controversial subject so far [45,46,47,48,49,50,51,52,53,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69]. The nature of the origin of the ferromagnetic properties of doped ZnO NPs has not been unambiguously described yet. Most researchers explain a change in the magnetic properties of Zn(1−x)MnxO NPs and Zn(1−y)CoyO NPs with the formation of, e.g., a secondary phase such as Co metal clusters, Co(OH)2, CoO, Co3O4, Co2O4, Mn2O3, Mn3O4, ZnMnO3 [48,49,50,51,52,53,54,55,56,57,58,59,60] and a spinel phase (ZnMn2O4, ZnCo2O4) [53,54,57,58,59,60,61,62,63,64], which are characterised by different magnetic properties (Table 1). However, a change in the magnetic properties of doped ZnO NPs, apart from a change in the level of the oxidation state of dopants and the presence of foreign phases, can also be explained by the presence of oxygen vacancies, the formation of dopant clusters, the exchange interaction between the local spin-polarised electrons and associated with specific defects and adsorbed elements at the surface of the NPs [66,67,68,69,70,71]. Martínez et al. [72], in turn, correlated the role of the microstructure with the change of the magnetic properties of Zn(1−y)CoyO NPs NPs, which demonstrates the complex nature of the magnetic properties of NPs. Based on the results of our own research [73], we believe that if the synthesis method employed permits the achievement of uniform substitution of Zn2+ ions in the ZnO crystalline lattice with dopant ions (Co2+ or Mn2+), such a material will be paramagnetic with some antiferromagnetic coupling. Currently, new methods of modifying ZnO properties are being sought to enable obtaining ZnO properties that satisfy the criteria of a DMS [46]. One of the solutions could be, e.g., codoping of ZnO with ions of Mn2+ and Co2+ (Zn(1−x−y)MnxCoyO). Codoping is a promising strategy that enables an effective modification of the magnetic and electronic properties of DMS materials. Research concerning codoping has been carried out for more than 20 years now [74]. It is believed that codoping might overcome the difficulties in bipolar doping and compensation in semiconductors [74]. Thanks to codoping, it is possible to increase the solubility of dopants, improve the stability of the expected defects, increase the carrier mobility and increase the activation rate by lowering the ionisation energy of acceptors and donors [74,75].
The relevant literature has reported a dozen or so methods of obtaining Zn(1−x−y)MnxCoyO nanostructures, an overview of which is presented in Table 2.
The magnetic properties of codoped ZnO depend chiefly on the method of obtaining it, the synthesis parameters and the sample preparation process (Table 2). Most authors modified the properties of the obtained Zn(1−x−y)MnxCoyO samples by selecting the temperature and gas atmosphere of the applied method or by soaking the samples additionally. An appropriate selection of the parameters for obtaining or preparing Zn(1−x−y)MnxCoyO samples permits the oxidation or reduction of dopant ions, which might result in the formation of a secondary phase of, e.g., dopant oxides and oxygen vacancies. A good example might be a study of Naeem et al. [99], where Zn(1−x−y)MnxCoyO NPs were obtained with the use of a chemical route and then were deliberately soaked at a temperature of 600 °C in an oxidising atmosphere and a reducing atmosphere. It was only Zn(1−x−y)MnxCoyO NP samples soaked in a reducing atmosphere that displayed ferromagnetic properties; their nature was explained by the formation of oxygen vacancies because Naeem et al. did not observe any foreign phase inclusions within the detection limit of the X-ray diffraction method. In order to understand the nature of the magnetic properties of Zn(1−x−y)MnxCoyO, the first-principles method based on the density functional theory is employed for their analysis [116,117,118].
Microwave solvothermal synthesis (MSS) has not been used for obtaining Zn(1−x−y)MnxCoyO NPs yet. Microwave technology is successfully applied for obtaining nanomaterials [119,120,121,122,123,124,125,126,127,128,129], including diverse nanostructures of ZnO [130,131,132,133,134,135,136,137,138,139,140,141,142]. The increased popularity of microwave-driven syntheses mainly arises from the decreasing prices of microwave reactors and the emergence of new reactor designs [143,144,145,146,147]. The main advantages of microwave heating in the case of ZnO NPs syntheses include:
-
reduction of the synthesis duration in comparison with conventional heating methods,
-
high homogeneity of NPs and narrow size distribution resulting from a small temperature gradient in the reaction vessel,
-
high product purity.
Our research conducted to date has shown that the microwave solvothermal synthesis of ZnO enables:
-
obtaining homogenous ZnO NPs with controlled size from circa 15 nm to 120 nm [140,141];
-
controlling the average size of ZnO NPs aggregates of 60, 90 and 120 nm at the same time preserving the constant size of single NPs, which are sized 27 nm [148];
-
obtaining homogenous ZnO NPs doped with Co2+ or Mn2+ ions [73,149,150];
-
obtaining ZnO NPs doped with Co2+ ions with controlled particle size between at least 28 nm and 53 nm [151].
The aim of this study is to prove that the microwave solvothermal synthesis enables obtaining homogeneous spherical Zn(1−x−y)MnxCoyO NPs with paramagnetic properties.

2. Materials and Methods

2.1. Substrates

Reagents used: zinc acetate dihydrate (Zn(CH3COO)2·2H2O, Zn(Ac)2·2H2O analytically pure, Chempur, Piekary Śląskie, Poland); manganese (II) acetate tetrahydrate (Mn(CH3COO)2∙4H2O, Mn(Ac)2·4H2O, analytically pure, Chempur, Piekary Śląskie, Poland), cobalt (II) acetate tetrahydrate (Co(CH3COO)2·4H2O, Co(Ac)2·4H2O pure, Sigma-Aldrich, St. Louis, MO, USA); ethylene glycol (EG, ethane-1,2-diol, C2H4(OH)2, pure, Chempur, Piekary Śląskie, Poland); deionised water (H2O) (specific conductance < 0.1 µS/cm, HLP 20UV, Hydrolab, Straszyn, Poland). The reagents were not subjected to any purification processes and were used as received. Only deionised water was used in the sample preparation process.

2.2. Synthesis of Zn(1−x−y)MnxCoyO NPs

Zn(1−x−y)MnxCoyO NPs were obtained with the use of the MSS method in accordance with the procedure described in our earlier publications [73,149,150,151,152]. Precursors of Zn(1−x−y)MnxCoyO were prepared by dissolving mixtures of crystalline cobalt acetate, manganese acetate, and zinc acetate in ethylene glycol. The composition of Zn(1−x−y)MnxCoyO NPs precursors summarised in Table 1 was calculated based on Equations (1) and (2), with the molar concentration of zinc acetate in ethylene glycol being constant and amounting to 0.3037 mol/dm3.
x Mn 2 + = n Mn 2 + n Zn 2 + + n Mn 2 + + n Co 2 +
y Co 2 + = n Co 2 + n Zn 2 + + n Mn 2 + + n Co 2 +
Precursor solutions (Table 3) were obtained by dissolving strictly defined compositions of mixtures of the acetates in glycol (150 mL) using a hot-plate magnetic stirrer (70 °C, 450 rpm, SLR, SI Analytics GmbH, Mainz, Germany). After the acetates dissolved, each solution was poured into a bottle (250 mL, PP), sealed and cooled down until it reached the ambient temperature.
The parameters of NPs syntheses carried out in a microwave reactor (Model 02-02600 W, 2.45 GHz, ERTEC, Poland) were as follows: volume of the precursor poured into a Teflon reaction container—75 mL, heating duration—25 min; maximum temperature of the bottom of the outer wall of the reaction vessel [140]—190 ± 5 °C; microwave power—100%, cool-down duration—20 min. The reaction duration was chosen based on our earlier research concerning the optimisation of parameters of the microwave solvothermal synthesis of ZnO NPs in ethylene glycol. Our primary assumption for the optimisation was to achieve the maximum reaction efficiency, while in the case of a reaction duration of 25 min the efficiency of zinc acetate conversion with identical synthesis parameters was circa 100% [141]. These setpoints of synthesis parameters were entered using the reactor’s control panel window (see the Supplementary Materials). An example profile of the course of the microwave solvothermal synthesis in the reactor model 02-02 is presented in Figure 1. The microwave heating in the employed reactor can be divided into the following stages:
  • Continuous microwave heating of the feedstock until the preset maximum temperature is reached (Tmax)—195 °C.
  • Once the feedstock reaches the maximum temperature of 195 °C, the reactor’s controller switches off the microwave heating and the feedstock temperature drops to the preset minimum temperature (Tmin)—185 °C.
  • Once the feedstock reaches Tmin 185 °C, the reactor’s controller switches on the continuous microwave heating and re-heats the sample to Tmax 195 °C.
  • Cyclic repetition of stages 2 and 3 until the preset reaction duration of 25 min is reached.
  • The reactor is cooled down by activating cold water flow in the metal body, inside which the Teflon reaction vessel containing the obtained post-reaction suspended matter is placed.
The synthesis of ZnO NPs codoped with Mn2+ and Co2+ ions in EG is presented by Equation (3):
Zn ( Ac ) 2 + Mn ( Ac ) 2 + Co ( Ac ) 2 C 2 H 4 ( OH ) 2 , H 2 O , T , P Zn ( 1 x y ) Mn x Co y O + esters + water .
Post-reaction suspended matter was centrifuged (MPW-350, MPW Med Instruments, Warsaw, Poland) and the supernatants were decanted. Afterwards, the sediments were rinsed intensively three times with water and centrifuged. Fifty millilitres of water were added to the obtained pastes; they were intensively stirred and then cooled down with liquid nitrogen and dried in a freeze dryer (Lyovac GT-2, SRK Systemtechnik GmbH, Riedstadt, Germany).

2.3. Characterisation Methods

The measurement procedures employed were described in our earlier publications [149,151]. The analysis of the phase composition was performed with the X-ray diffraction method (XRD) (2 theta from 10° to 150°, room temperature, CuKα1, X’Pert PRO, Panalytical, The Netherlands). The parameters of the crystalline lattice, a and c, were determined by the Rietveld method in Fityk software, version 0.9.8. Crystallite size was calculated by means of Scherrer’s formula [149] and an analysis of the profile of XRD peaks with the FW15/45M method [153,154,155], which also enabled the calculation of the crystallite size distribution.
The morphology of NPs was defined using scanning electron microscopy (SEM) (ZEISS, Ultra Plus, Oberkochen, Germany).
Density was measured using a helium pycnometer at a temperature of 24 ± 1 °C (ISO 12154:2014, AccuPyc II 1340 FoamPyc V1.06, Micromeritics, Norcross, GA, USA).
The specific surface area of NPs was determined through an analysis of nitrogen adsorption isotherm by the BET (Brunauer-Emmett-Teller) method (ISO 9277:2010, Gemini 2360, V 2.01, Micromeritics). Based on the results of the specific surface area and density, the average particle size was calculated with the assumption that all particles were spherical and identical [149].
The chemical composition of the samples dissolved in nitric acid (V) was analysed with the use of argon inductively coupled plasma optical emission spectrometry (ICP-OES) (Thermo Fisher Scientific, iCAP model 6000, Waltham, MA, USA) [73]. The quantitative microanalysis of the Zn, Mn and Co content in the pressed NP samples was performed using energy-dispersive spectrometry (EDS) (Quantax 400, Bruker, Billerica, MA, USA) [151].
The colour analysis by means of the Red‒Green‒Blue (RGB; value: 0 to 1023) and Hue‒Saturation‒Luminance (HSL; value: 0 to 1000) colour model was carried out with the RGB-2000 metric by VOLTCRAFT (Conrad Electronic SE, Wernberg-Köblitz, Germany) in accordance with the manufacturer’s recommendations.
The magnetisation measurements are performed with the use of a SQUID-type magnetometer (liquid helium cooled MPMSXL device manufactured by Quantum Design, Inc., San Diego, CA, USA) in the temperature range 2–300 K and magnetic fields up to 7 T.

3. Results and Discussion

3.1. Morphology

Figure 2 and Figure 3 present selected representative SEM images of Zn(1− x−y)MnxCoyO NP samples. An impact of the content of dopants on the morphology of Zn(1−x−y)MnxCoyO is noticeable. Powders of ZnO and Zn(0.98)Mn0.01Co0.01O were composed of loose homogeneous spherical particles sized 20–50 nm. Powders of Zn0.90Co0.05Mn0.05O and Zn0.8Co0.1Mn0.1O, in turn, were composed of compact homogeneous spherical NPs sized 20–40 nm, which formed conglomerates sized between 1 µm and 3 µm resembling a “cauliflower” in terms of shape and structure. A similar impact of a dopant on a change in NPs morphology was observed in our earlier studies on doped ZnO [149,151]. In order to eliminate the effect of aggregation of the obtained Zn(1−x−y)MnxCoyO NPs, the synthesis parameters must be individually optimised for each composition of Zn(1−x−y)MnxCoyO NPs. We have shown that microwave solvothermal synthesis permits controlling the size of ZnO NPs aggregates through a change of the microwave power used for ZnO NPs synthesis [148]. Zhang et al. [156] described the impact of a change in the solvothermal synthesis temperature on the size of ZnO NPs aggregates.
Figure 3 presents the morphology of Zn0.70Co0.15Mn0.15O powder. SEM images show two products of synthesis: ZnO NPs aggregates, and a lamellar structure, which might be an unreacted intermediate in the form of hydroxide metal acetates with the general formula of M5(OH)(10−z)(CH3COO)Z·nH2O, where M = (Zn, Co and Mn) [157], or a compound of codoped hydroxide metal acetates Zn5(1−x−y)Mn(5x)Co(5y)(OH)(10−z)(CH3COO)z·nH2O [141,151]. Information regarding the phase composition of the Zn0.70Co0.15Mn0.15O sample will be provided by XRD. The product prevalent in the obtained Zn0.70Co0.15Mn0.15O sample is NPs, which is illustrated by the SEM image in Figure 3c.

3.2. Phase Composition and Lattice Parameters

Despite the presence of the lamellar structure visible in SEM images in the Zn0.7Co0.15Mn0.15O sample, all diffraction peaks in Figure 4 were attributed exclusively to the hexagonal phase ZnO (JCPDS No. 36-1451). The XRD test did not show the presence of the secondary phase, which could be M5(OH)(10−z)(CH3COO)Z·nH2O (lamellar structure) [157].
The hexagonal phase of M5(OH)(10−z)(CH3COO)Z·nH2O has a diffraction peak below 10° (2 theta angle) [141]. In order to verify the presence of a foreign phase in the Zn0.7Co0.15Mn0.15O sample, XRD measurement was performed again with the range from 5° to 100° (Figure 5). All diffraction peaks in Figure 5 were attributed exclusively to the hexagonal phase ZnO (JCPDS No. 36-1451). The absence of the secondary phase (Figure 3) in XRD results may mean that:
-
the amount of the secondary phase was below the detection limit of the XRD method,
-
the secondary phase was an amorphous material,
-
the secondary phase is lamellar-shaped ZnO.
At this stage of the research, we are unable to identify what the lamellar structure product in the Zn0.7Co0.15Mn0.15O sample is (Figure 3) and we do not know whether its presence can be eliminated through optimisation of microwave solvothermal synthesis processes. This requires further research.
Theorem-type environments (including propositions, lemmas, corollaries, etc.) can be formatted as follows: ZnO is characterised by the hexagonal wurtzite structure (295 K, h-ZnO space group P63mc) with the lattice parameters of a = 3.2498 Å and c = 5.2066 Å [6], while the c/a lattice parameter ratio in ZnO is 1.6021 and is close to the c/a value = 1.6330 for the theoretical close-packed hexagonal structure (hcp) [151,158]. The ionic radius of Zn2+ is 0.74 Å [159]. Manganese oxide (II) MnO has a cubic rock salt structure (295 K, c-MnO, space group Oh5—Fm3m) with the lattice parameter of a = 4.4475 Å [160], and the ionic radius of Mn2+ is 0.83 Å [161]. Cobalt oxide (II) CoO crystallises in two crystalline phases: cubic rock salt CoO (295 K, c-CoO, space group Fm3m) with the lattice parameter of a = 4.2581 Å, and hexagonal wurtzite CoO (77 K, h-CoO, space group P63mc) with lattice parameters of a = 5.183 Å i c = 3.017 Å [160]. The ionic radius of Co2+ is 0.745 Å [162].
The results of our earlier studies [73,149] proved that the lattice parameters a and c in ZnO change in line with the change in Co2+ or Mn2+ ion content, which we explained with the impact of changes in the values of ion radii of dopants in relation to Zn2+. However, each dopant gave rise to a distinct change in the value of the lattice parameters of doped ZnO depending on the content and type. For Co2+ dopant ions, whose ion radius is almost identical to that of Zn2+, the value of a lattice parameter in Zn(1−y)CoyO samples grew in line with the increase in the dopant content from 1% to 15% [73]. The value of the c lattice parameter, in turn, grew when the nominal dopant content ranged from 1% to 5%, and dropped when the content was 5 to 15%. In the case of the dopant of Mn2+ ions, where their ionic radius is greater than Zn2+ radius by as much as 0.09 Å, both lattice parameters, a and c, in Zn(1−x)MnxO samples increased their values within the dopant content range of 1 to 25 mol % [149].
The calculated a and c lattice parameters of Zn(1−x−y)MnxCoyO samples are presented in Table 4 and in Figure 6. The value of both lattice parameters a and c grew in line with the increase in the content of Mn2+ and Co2+ dopants in ZnO (Figure 6). Table 4 also presents the lattice parameters of Zn0.85Mn0.15O NPs and Zn0.85Co0.15O NPs as comparative samples for the Zn0.70Mn0.15Co0.15O sample. According to our assumptions, the principal impact on the value of ZnO lattice parameters was exerted by the dopant with the greater ionic radius since a and c lattice parameters for the Zn0.70Mn0.15Co0.15O sample are comparable to the values for the Zn0.85Mn0.15O sample. The change in a and c lattice parameters in the obtained Zn(1−x−y)MnxCoyO samples proves an effective substitution of Zn2+ ions with Mn2+ and Co2+ dopant ions in the ZnO crystalline lattice. Similar dependencies on changes to Mn‒Co codoped ZnO lattice parameters were described by Adbullahi et al. [105].

3.3. Impact of Chemical Composition on the Colour of Codoped ZnO NPs

The results of the dopant content analysis are presented in Table 5. The differences between the results of the analysis carried out by the EDS method and the ICP-OES method arise from the sensitivity, accuracy and limitations of these analytical techniques as regards the quantitative determination of manganese, cobalt and zinc. The NP samples for the EDS analysis were pressed to form pastilles so that the sample surface was flat and smooth and the porosity of the examined material was reduced as much as possible (ISO 22309:2011). The ICP method is considered a source of the most representative results owing to its numerous advantages, e.g., low detection limits, a wide linear dynamic range, high precision [163]. Our interesting findings include efficiencies of doping for the obtained samples (Table 6). In line with the increase in the nominal contents of two dopants in precursor solutions, the efficiency of doping with Mn2+ ions decreased and at the same time the efficiency of doping with Co2+ ions in codoped ZnO NPs increased. The achieved low efficiency of doping with Mn2+, which was between 16% and 25%, can be explained by the difference in the sizes of the ionic radii of Mn2+ and Zn2+ being 0.09 Å. The efficiency of doping with Co2+ ions, in turn, was as many as 3–4 times greater than the efficiency of doping with Mn2+ ions, which is explained by the negligible difference in the values of Co2+ and Zn2+ ionic radii. An advantage of the method we have developed is the use of a solution of salts dissolved in ethylene glycol, which enables obtaining homogenous codoped NPs. The mechanism of the MSS reaction of Zn(1−x−y)MnxCoyO NPs permits the substitution of Zn2+ ions with the optimum number of Mn2+ and Co2+ ions in the crystalline lattice of the ZnO being formed, depending on the precursor composition. The remaining quantity of Mn2+ and Co2+, which has not been integrated into the crystalline lattice of ZnO, in turn, remains in the form of unreacted acetate salts dissolved in ethylene glycol [73]. In the sol-gel co-precipitation method, the codoping efficiency is 100% [102], which considerably reduces the costs of synthesis and chemical waste disposal. Nevertheless, thanks to the employment of ethylene glycol, a solvent with poor reducing properties, the MSS method prevents a change in the oxidation state of Mn2+ and Co2+ dopants [73,149,151], which considerably limits the possibility of formation of foreign phase inclusions. Another advantage of the MSS method is the use of a low synthesis temperature (190–220 °C) and a short synthesis time compared to other methods enumerated in Table 2, which prevents uncontrolled NPs growth, formation of NPs aggregates as a consequence of sintering, and formation of foreign phase inclusions during NPs growth.
It is generally known that ZnO codoping with Mn2+ and Co2+ ions causes a change in the optical properties and a reduction of the band gap width [103,105]. Figure 7 and Figure 8 present a visual comparison of changes in the colours of Zn(1−x−y)MnxCoyO samples depending on the content of dopants, while Table 7 summarises the results of the analysis of colours by the RGB and HSL colour models. ZnO codoping with Mn2+ and Co2+ ions resulted in a change of sample colour from light green to dark olive green. A similar change of Zn(1−x−y)MnxCoyO powder colours is reported by Yin-Hua et al. [100]. The colour of the codoped ZnO is a resultant of the impact of the presence of two dopants in the ZnO crystalline lattice. Our earlier studies [149,150] showed that ZnO doped with Mn2+ ions changes colour from yellow to orange (Figure 7c) when the quantity of the Mn2+ dopant increases, while Co-doped ZnO NPs are green and the intensity of the colour depends on the content of the Co2+ dopant (Figure 7b). Based on the impact of a single dopant on the colour of ZnO, it can be inferred that the green colour of Zn(1−x−y)MnxCoyO is attributed mainly to the presence of the Co2+ dopant. The modification of the colour of ZnO NPs through their codoping permits the application of Zn(1−x−y)MnxCoyO nanopowders as colourful pigments. Obviously, when using Zn(1−x−y)MnxCoyO NPs as pigments, their applications should be selected such that their potential toxicity is irrelevant.

3.4. Density, Specific Surface Area and Size Distribution of NP

The obtained results are summarised in Table 8. The density of an undoped ZnO NP sample is 5.25 g/cm3 and is smaller than the value of 5.61 g/cm3 for the theoretical density of ZnO [164]. The difference between theoretical densities of the materials and real densities of these materials in the nano form is well known [140,151,165,166] and arises mainly from the imperfection of the real crystalline structure of nanomaterial surface, deviations from the stoichiometric composition, or the presence of hydroxides. With the theoretical density of CoO (6.45 g/cm3) and MnO (5.37 g/cm3) taken into consideration, it could be assumed that the density of the samples would increase in line with the growth of the dopant content in Zn(1−x−y)MnxCoyO. However, the density of Zn(1−x−y)MnxCoyO samples dropped from 5.25 g/cm3 to 5.06 g/cm3 in line with the increase in the dopant content, which could be explained by three causes:
-
a different volume and packing of the unit cell of ZnO than in the case of MnO and CoO [6];
-
lower atomic mass of the dopants (Co2+—≈58.93 u; Mn2+—≈54.94 u) in comparison to the substituted Zn2+ (≈65.38 u) atoms in the ZnO crystalline structure;
-
increasing number of defects in the ZnO crystalline structure resulting from the growth of the content of Co2+ and Mn2+ dopants.
In line with the increase in the dopant content in Zn(1−x−y)MnxCoyO (Table 8), the specific surface area of the samples increased from 39.8 m2/g to 56.4 m2/g, while the average particle size calculated based on the specific surface area and density decreased from 29 nm to 21 nm. The average crystallite size calculated by the two methods fell within the range of standard deviations of these results. The ratio of the sizes of dc i da crystallites proves a change in the asymmetry of the obtained Zn(1−x−y)MnxCoyO crystallites, and similar dc/da values were obtained for Zn(1−x)MnxO NPs depending on the dopant content [149]. The obtained crystallite size distribution of the Zn(1−x−y)MnxCoyO samples showed that the growth of the dopant content resulted in a wider distribution (Figure 9). The differences in the results between the average particle size and the average crystallite size might be caused by the adopted assumptions of the calculation methods employed [149,151,153,167], and the similar values of average particle and crystallite sizes prove that Zn(1−x−y)MnxCoyO particles are made of single crystallites.

3.5. Magnetic Characterisation of the Zn(1−x−y)MnxCoyO NPs

The representative magnetisation data are shown in Figure 10, where magnetisation as measured versus magnetic field is depicted at T = 2 K (a) and T = 300 K (b). Low-temperature data overall show typical paramagnetic behaviour: magnetisation increases monotonously in line with the increasing magnetic field and tends to saturate for the highest fields. For the samples with low concentrations of magnetic ions, saturation is nearly perfect, whereas for the highest concentrations magnetisation does not saturate in the applied field range. Such behaviour is typical for a paramagnetic system of localised magnetic moments of transition metals (TM) ions, coupled by an antiferromagnetic exchange interaction and was commonly observed for different. For lightly Mn2+/Co2+-doped samples this picture is confirmed by high-temperature data (T = 300 K), where magnetisation is a linear function of the magnetic field (cf. Figure 10b). On the other hand, high-temperature data for high Mn2+/Co2+ concentrations reveal a different behaviour: magnetisation rises fast at low fields (B < 1 T), with a tendency to saturation, and then shows a linear field dependence (for B > 1 T). This suggests two components of the measured magnetisation: one purely paramagnetic, responsible for the linear field dependence and the second originating from ferro/ferrimagnetic phase arising during the growth process of the samples (e.g., Co3O4, Mn3O4). We note that such a situation was widely encountered for DMS synthesised with TM concentration close to the solubility limits [168,169,170].
Having the above in mind, the measured magnetic moment of each sample can be regarded as a sum of paramagnetic contributions of localised magnetic moments of Mn2+ or Co+2 ions, diamagnetic contributions of the NP lattice and the sample holder (parafilm wrapper, glue), as well as possible contributions of unintentional impurities/secondary phases. Therefore, the measured magnetic moment can be expressed in the form:
Mexp(B, T) = MNP(B, T) + Xdia*B + C,
where MNP(B, T) is the total magnetisation of the magnetic moments of NPs, Xdia is the sum of diamagnetic susceptibility of NP, parafilm and glue (assumed to be temperature independent in the studied temperature range) and C represents the contribution from possible ferro/ferrimagnetic phases, as well as ferro/ferrimagnetic contributions of the ingredients used in the synthesis of the samples, parafilm and glue. As may be noticed from Figure 10b, the diamagnetic contribution for low concentrations of Mn/Co is sizeable and dominates the measured magnetic moment at high temperatures, e.g., at T = 300 K the measured magnetic moment is diamagnetic. In such a case, precise values of Xdia and C for the sample in question are crucial. Instead of applying a standard way to evaluate Xdia, i.e., measuring the undoped sample (ZnO), it is proposed to eliminate Xdia*B + C contributions by subtracting the high-temperature magnetic moment (where the paramagnetic contribution of Mn/Co magnetic moments is largely quenched) from the low-temperature one. Assuming that Xdia and C are temperature-independent (which is true to a large extent), the following is obtained:
Mexp(B, T = 2 K) − Mexp(B, T = 300 K) = MNP(B, T = 2 K) − MNP(B, T = 300 K),
Given the previous results for Mn2+- and Co2+-base DMS, MNP(B, T) can be assumed in the following form:
MNP(B, T) = A g μB S BS(B, T),
where BS(B, T) is the Brillouin function for spin S = 5/2 (Mn2+) or S = 3/2 (Co2+), g = 2.00 is the g-factor, μB is the Bohr magneton and A is the number of spins (magnetic moments) in the sample. A possible interaction between spins can be taken into account by assuming effective temperature Teff = T − T0, instead of experimental temperature T [171,172]. We recall that T0 < 0 corresponds to antiferromagnetic (AFM) interactions, while T0 > 0 means ferromagnetic (FM) interactions.
In order to demonstrate how the proposed method works, we start with NPs doped only with one type of ions, i.e., Mn2+ or Co2+. The results concerning the characterisation of Zn0.99Mn0.01O, Zn0.85Mn0.15O, Zn0.99Co0.01O and Zn0.85Co0.15O samples are included in the Supplementary Materials Figure 11a shows Mexp(B, T = 2 K)—Mexp(B, T = 300 K), as well as the fit with Equations (5) and (6), where A was the only adjustable parameter (spin S = 5/2 and T0 = 0 were fixed).
The slower saturation of magnetisation MNP(B, T) than the Brillouin function reflects AFM interaction between Mn2+ ions, which is expected for the system with about 1% of Mn2+ ions [172]. As mentioned above, this effect can be taken into account by introducing effective temperature Teff = T − T0 and considering T0 as an adjustable parameter. Figure 11b shows Mexp(B, T = 2 K)—Mexp(B, T = 300 K), as well as the fits with Equations (5) and (6), where A and T0 are adjustable parameters (spin S = 5/2 or S = 3/2 are fixed).
Satisfactory matching was obtained for all the samples. Effective temperatures T0 are all negative, indicating AFM exchange interactions between Mn2+ and Co2+ ions, as expected [171,172].
For the samples codoped with both Mn2+ and Co2+ the following function was used:
MNP(B, T) = AMn g μB SMn BS(B, Teff) + ACo g μB SCo BS(B, Teff),
where AMn (ACo) corresponds to number of Mn2+ (Co2+) ions, SMn = 5/2, SCo = 3/2. In order to avoid a large number of fitting parameters only one effective temperature was used, which means that T0 is a common parameter for all interactions in the NP, i.e., Mn2+-Mn2+, Co2+-Co2+ and Mn2+-Co2+.
The results of the fittings are shown in Figure 12. For all the samples T0 is negative, suggesting preferred antiferromagnetic interactions between Mn and Co ions.
To summarise the magnetic properties, it can be stated that the discussed NP samples can be generally described as the systems of localised magnetic moments arising from Mn2+ or Co2+ d-shell electrons located at Zn2+ lattice sites. The paramagnetic properties of these systems can be reasonably well described by the effective Brillouin function, with an indication of AFM exchange interactions between Mn2+ or Co2+ ions. Moreover, for the Zn0.70Mn0.15Co0.15O samples with the highest TM ions concentrations, an additional ferromagnetic-type magnetic phase is observed, most probably originating from crystalline phases other than Zn(1−x−y)MnxCoyO.
The Zn0.70Mn0.15Co0.15O NP sample is a good example, confirming our earlier research results [73]. Namely, we stated that if a ZnO sample displays ferromagnetic properties, this may be caused by the presence of a foreign phase or a change in the dopant oxidation state. XRD tests did not indicate the presence of a foreign phase in the Zn0.70Mn0.15Co0.15O NP sample, but the SEM images (Figure 3a) show a small quantity of a lamellar structure, which probably is a foreign phase.

4. Conclusions

ZnO nanoparticles strongly codoped with Mn2+ and Co2+ ions up to the nominal content of 15 mol % were obtained for the first time by means of the microwave solvothermal synthesis method. The reaction precursors were solutions of zinc acetate, manganese acetate (II), and cobalt acetate (II) dissolved in ethylene glycol. No other phases were detected by means of X-ray diffraction in all obtained samples. The actual Mn2+ and Co2+ ions dopant content ranged from 15.67% to 25.00% and from 69.00% and 80.47% of the nominal one, respectively. The colour of Zn(1−x−y)MnxCoyO NP samples changed in line with the dopants value from white (undoped ZnO), through light green (Zn0.98Mn0.01Co0.01O), to dark olive green (Zn0.70Co0.15Mn0.15O). The obtained results of the chemical composition, the increase in the crystalline lattice parameters, and the changes of colours confirmed the successful codoping of ZnO NPs with Mn2+ and Co2+ ions. Zn(1−x−y)MnxCoyO powders are composed of homogenous spherical NPs, which, depending on the dopants content, form aggregates resembling a “cauliflower” structure. Only in the Zn0.7Mn0.15Co0.15O sample was a small quantity of an unidentified lamellar structure observed. The growth of the dopant content from 0% (x = y) to 15% (x = y) in Zn(1−x−y)MnxCoyO NPs caused an increase in the specific surface area from 39.8 m2/g to 56.4 m2/g; a decrease in density from 5.25 g/cm3 to 5.06 g/cm3; and a decrease in the average NPs size from 29 nm to 21 nm.
The magnetic properties of the nanoparticles can be described as systems of localised magnetic moments arising from Mn2+ or Co2+ d-shell electrons located at Zn2+ lattice sites. The paramagnetic properties of these nanoparticles are described by the effective Brillouin function, with an indication of antiferromagnetic exchange interactions between Mn2+ or Co2+ ions. For the Zn0.70Mn0.15Co0.15O samples with the highest ion concentrations, an additional, ferromagnetic-type magnetic phase is observed.
The achieved results confirm the numerous advantages of the microwave solvothermal synthesis in obtaining codoped ZnO NPs.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4352/8/11/410/s1.

Author Contributions

J.W., W.L., M.O. and J.S. conceived and designed the experiments; J.W., M.O. and S.G. performed the experiments; J.W., M.O., J.S., T.C., S.G., A.M., A.T. and W.L. analysed the data; J.W., J.S., A.T. and W.L. contributed reagents/materials/analysis tools; J.W., M.O., J.S., A.T. and W.L. wrote the paper.

Funding

The work was supported by the Institute of High Pressure Physics, the Polish Academy of Sciences.

Acknowledgments

The research was carried out using equipment funded by the CePT project, reference: POIG.02.02.00-14-024/08, financed by the European Regional development Fund within the Operational Programme “Innovative Economy” for 2007–2013. The authors would also like to thank Magdalena Grochowska (Faculty of Physics, University of Warsaw), Jan Mizeracki (Institute of High Pressure Physics of the Polish Academy of Sciences) and Roman Mukhovskyi.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Fortunato, E.; Gonçalves, A.; Pimentel, A.; Barquinha, P.; Gonçalves, G.; Pereira, L.; Ferreira, I.; Martins, R. Zinc oxide, a multifunctional material: From material to device applications. Appl. Phys. A 2009, 96, 197–205. [Google Scholar] [CrossRef]
  2. Ali, A.; Phull, A.R.; Zia, M. Elemental Zinc to Zinc nanoparticles: Is ZnO NPs crucial for life? Synthesis, toxicological and environmental concerns. Nanotechnol. Rev. 2018, 7, 413–441. [Google Scholar] [CrossRef]
  3. Ashrafi, A.; Jagadish, C. Review of zincblende ZnO: Stability of metastable ZnO phases. J. Appl. Phys. 2007, 102, 071101. [Google Scholar] [CrossRef]
  4. Ozgur, U.; Alivov, Y.I.; Liu, C.; Teke, A.; Reshchikov, M.A.; Dogan, S.; Avrutin, V.; Cho, S.J.; Morkoc, H. A comprehensive review of ZnO materials and devices. J. Appl. Phys. 2005, 98, 041301. [Google Scholar] [CrossRef]
  5. Jagadish, C.; Pearton, S. Zinc Oxide Bulk, Thin Films and Nanostructures, 1st ed.; Elsevier: Oxford, UK, 2006; ISBN 978-0-08-044722-3. [Google Scholar]
  6. Morkoç, H.; Özgür, Ü. Zinc Oxide: Fundamentals, Materials and Device Technology, 1st ed.; WILEY-VCH: Weinheim, Germany, 2009; ISBN 978-3-527-40813-9. [Google Scholar]
  7. Ozgur, U.; Hofstetter, D.; Morkoc, H. ZnO devices and applications: A review of current status and future prospects. Proc. IEEE 2010, 98, 1255–1268. [Google Scholar] [CrossRef]
  8. Litton, C.W.; Reynolds, D.C.; Collins, T.C. Zinc Oxide Materials for Electronic and Optoelectronic Device Applications, 1st ed.; John Wiley & Sons, Ltd.: Chichester, UK, 2011; ISBN 9780470519714. [Google Scholar]
  9. Willander, M. Zinc Oxide Nanostructures: Advances and Applications, 1st ed.; Pan Stanford: New York, NY, USA, 2014; ISBN 9789814411332. [Google Scholar]
  10. Djurišić, A.B.; Ng, A.M.C.; Chen, X.Y. ZnO Nanostructures for Optoelectronics: Material Properties and Device Applications. Prog. Quantum Electron. 2010, 34, 191–259. [Google Scholar] [CrossRef]
  11. Oprea, O.; Andronescu, E.; Ficai, D.; Ficai, A.; Oktar, F.N.; Yetmez, M. ZnO Applications and Challenges. Curr. Org. Chem. 2014, 18, 192–203. [Google Scholar] [CrossRef]
  12. Martínez-Carmona, M.; Gun’ko, Y.; Vallet-Regí, M. ZnO Nanostructures for Drug Delivery and Theranostic Applications. Nanomaterials 2018, 8, 268. [Google Scholar] [CrossRef] [PubMed]
  13. Jiang, J.; Pi, J.; Cai, J. The Advancing of Zinc Oxide Nanoparticles for Biomedical Applications. Bioinorg. Chem. Appl. 2018, 2018, 1062562. [Google Scholar] [CrossRef] [PubMed]
  14. Sirelkhatim, A.; Mahmud, S.; Seeni, S.; Kaus, A.S.N.; Ann, L.C.; Bakhori, S.K.M.; Hasan, H.; Mohamad, D. Review on Zinc Oxide Nanoparticles: Antibacterial Activity and Toxicity Mechanism. Nano-Micro Lett. 2015, 7, 219–242. [Google Scholar] [CrossRef] [Green Version]
  15. Zhang, Y.; Nayak, T.R.; Hong, H.; Cai, W. Biomedical Applications of Zinc Oxide Nanomaterials. Curr. Mol. Med. 2013, 13, 1633–1645. [Google Scholar] [CrossRef] [PubMed]
  16. Sruthi, S.; Mohanan, P.V. Engineered Zinc Oxide Nanoparticles; Biological Interactions at the Organ Level. Curr. Med. Chem. 2016, 23, 4057–4068. [Google Scholar] [CrossRef] [PubMed]
  17. Kalpana, V.N.; Devi Rajeswari, V. A Review on Green Synthesis, Biomedical Applications, and Toxicity Studies of ZnO NPs. Bioinorg. Chem. Appl. 2018, 2018, 3569758. [Google Scholar] [CrossRef] [PubMed]
  18. Cierech, M.; Kolenda, A.; Grudniak, A.M.; Wojnarowicz, J.; Woźniak, B.; Gołaś, M.; Swoboda-Kopeć, E.; Łojkowski, W.; Mierzwińska-Nastalska, E. Significance of polymethylmethacrylate (PMMA) modification by zinc oxide nanoparticles for fungal biofilm formation. Int. J. Pharm. 2016, 510, 323–335. [Google Scholar] [CrossRef] [PubMed]
  19. Cierech, M.; Wojnarowicz, J.; Szmigiel, D.; Bączkowski, B.; Grudniak, A.; Wolska, K.; Łojkowski, W.; Mierzwińska-Nastalska, E. Preparation and characterization of ZnO-PMMA resin nanocomposites for denture bases. Acta Bioeng. Biomech. 2016, 18, 31–41. [Google Scholar] [CrossRef] [PubMed]
  20. Cierech, M.; Osica, I.; Kolenda, A.; Wojnarowicz, J.; Szmigiel, D.; Łojkowski, W.; Kurzydłowski, K.; Ariga, K.; Mierzwienska-Nastalska, E. Mechanical and Physicochemical Properties of Newly Formed ZnO-PMMA Nanocomposites for Denture Bases. Nanomaterials 2018, 8, 305. [Google Scholar] [CrossRef] [PubMed]
  21. Paszek, E.; Czyz, J.; Woźniacka, O.; Jakubiak, D.; Wojnarowicz, J.; Łojkowski, W.; Stępień, E. Zinc oxide nanoparticles impair the integrity of human umbilical vein endothelial cell monolayer in vitro. J. Biomed. Nanotechnol. 2012, 8, 957–967. [Google Scholar] [CrossRef] [PubMed]
  22. Pokrowiecki, R.; Pałka, K.; Mielczarek, A. Nanomaterials in dentistry: A cornerstone or a black box? Nanomedicine 2018, 13, 639–667. [Google Scholar] [CrossRef] [PubMed]
  23. Jamalullail, N.; Mohamad, I.S.; Norizan, M.N.; Mahmed, N.; Taib, B.N. Recent improvements on TiO2 and ZnO nanostructure photoanode for dye sensitized solar cells: A brief review. EPJ Web Conf. 2017, 162, 01045. [Google Scholar] [CrossRef]
  24. Luo, J.; Wang, Y.; Zhang, Q. Progress in perovskite solar cells based on ZnO nanostructures. Sol. Energy 2018, 163, 289–306. [Google Scholar] [CrossRef]
  25. Melo, A.H.N.; Macêdo, M.A. Permanent Data Storage in ZnO Thin Films by Filamentary Resistive Switching. PLoS ONE 2016, 11, e0168515. [Google Scholar] [CrossRef] [PubMed]
  26. Omelchenko, M.M.; Wojnarowicz, J.; Salamonczyk, M.; Lojkowski, W. Lyotropic liquid crystal based on zinc oxide nanoparticles obtained by microwave solvothermal synthesis. Mater. Chem. Phys. 2017, 192, 383–391. [Google Scholar] [CrossRef]
  27. Salzano de Luna, M.; Galizia, M.; Wojnarowicz, J.; Rosa, R.; Lojkowski, W.; Acierno, D.; Filippone, G.; Leonelli, C. Dispersing hydrophilic nanoparticles in hydrophobic polymers: HDPE/ZnO nanocomposites by a novel template-based approach. Express Polym. Lett. 2014, 8, 362–372. [Google Scholar] [CrossRef] [Green Version]
  28. Huang, J.; Huang, G.; An, C.; He, Y.; Yao, Y.; Zhang, P.; Shen, J. Performance of ceramic disk filter coated with nano ZnO for removing Escherichia coli from water in small rural and remote communities of developing regions. Environ. Pollut. 2018, 238, 52–62. [Google Scholar] [CrossRef] [PubMed]
  29. Ong, C.B.; Ng, L.Y.; Mohammad, A.W. A review of ZnO nanoparticles as solar photocatalysts: Synthesis, mechanisms and applications. Renew. Sustain. Energy Rev. 2018, 81, 536–551. [Google Scholar] [CrossRef]
  30. Hamid, S.B.A.; Teh, S.J.; Lai, C.W. Photocatalytic Water Oxidation on ZnO: A Review. Catalysts 2017, 7, 93. [Google Scholar] [CrossRef]
  31. Tudose, I.V.; Suchea, M. ZnO for photocatalytic air purification applications. IOP Conf. Ser. Mater. Sci. Eng. 2016, 133, 012040. [Google Scholar] [CrossRef] [Green Version]
  32. Djurišić, A.B.; Chen, X.; Leung, Y.H.; Ng, A.M.C. ZnO nanostructures: Growth, properties and applications. J. Mater. Chem. 2012, 22, 6526–6535. [Google Scholar] [CrossRef]
  33. Chaudhary, S.; Umar, A. ZnO Nanostructures and Their Sensing Applications: A Review. Nanosci. Nanotechnol. Lett. 2017, 9, 1787–1826. [Google Scholar] [CrossRef]
  34. Woźniak, B.; Dąbrowska, S.; Wojnarowicz, J.; Chudoba, T.; Łojkowski, W. Content available remote Coating synthetic materials with zinc oxide nanoparticles acting as a UV filter. Glass Ceram. 2017, 3, 15–17. [Google Scholar]
  35. Fu, Y.; Fu, W.; Liu, Y.; Zhang, G.; Liu, Y.; Yu, H. Comparison of ZnO nanorod array coatings on wood and their UV prevention effects obtained by microwave-assisted hydrothermal and conventional hydrothermal synthesis. Holzforschung 2015, 69, 1009–1014. [Google Scholar] [CrossRef]
  36. Wallenhorst, L.; Gurău, L.; Gellerich, A.; Militz, H.; Ohms, G.; Viöl, W. UV-blocking properties of Zn/ZnO coatings on wood deposited by cold plasma spraying at atmospheric pressure. Appl. Surf. Sci. 2018, 434, 1183–1192. [Google Scholar] [CrossRef]
  37. Sabir, S.; Arshad, M.; Chaudhari, K.C. Zinc Oxide Nanoparticles for Revolutionizing Agriculture: Synthesis and Applications. Sci. World J. 2014, 2014, 925494. [Google Scholar] [CrossRef] [PubMed]
  38. Singh, A.; Singh, N.B.; Afzal, S.; Singh, T.; Hussain, I. Zinc oxide nanoparticles: A review of their biological synthesis, antimicrobial activity, uptake, translocation and biotransformation in plants. J. Mater. Sci. 2018, 53, 185–201. [Google Scholar] [CrossRef]
  39. Pearton, S.J.; Norton, D.P.; Heo, Y.W.; Tien, L.C.; Ivill, M.P.; Li, Y.; Kang, B.S.; Ren, F.; Kelly, J.; Hebard, A.F. ZnO spintronics and nanowire devices. J. Electron. Mater. 2006, 35, 862–868. [Google Scholar] [CrossRef]
  40. Lu, J.W.; Chen, E.; Kabir, M.; Stan, M.R.; Wolf, S.A. Spintronics technology: Past, present and future. Int. Mater. Rev. 2016, 61, 456–472. [Google Scholar] [CrossRef]
  41. Awschalom, D.D.; Flatté, M.E. Challenges for semiconductor spintronics. Nat. Phys. 2007, 3, 153–159. [Google Scholar] [CrossRef]
  42. Dietl, T.; Ohno, H.; Matsukura, M.; Cibert, J.; Ferrand, D. Zener Model Description of Ferromagnetism in Zinc-Blende Magnetic Semiconductors. Science 2000, 287, 1019–2000. [Google Scholar] [CrossRef] [PubMed]
  43. Sato, K.; Katayama-Yosida, H. First principles materials design for semiconductor spintronics. Semicond. Sci. Technol. 2002, 17, 367–376. [Google Scholar] [CrossRef]
  44. Mustaqima, M.; Liu, C. ZnO-based nanostructures for diluted magnetic semiconductor. Turk. J. Phys. 2014, 38, 429–441. [Google Scholar] [CrossRef] [Green Version]
  45. Djerdj, I.; Jaglicić, Z.; Arcon, D.; Niederberger, M. Co-Doped ZnO nanoparticles: Minireview. Nanoscale 2010, 2, 1096–1104. [Google Scholar] [CrossRef] [PubMed]
  46. Yang, Z. A perspective of recent progress in ZnO diluted magnetic semiconductors. Appl. Phys. A 2013, 112, 241–254. [Google Scholar] [CrossRef]
  47. Liu, C.; Yun, F.; Morkoç, H. Ferromagnetism of ZnO and GaN: A Review. J. Mater. Sci. Mater. Electron. 2005, 16, 555–597. [Google Scholar] [CrossRef]
  48. Naeem, M.; Hasanain, S.K.; Kobayashi, M.; Ishida, Y.; Fujimori, A.; Buzby, S.; Shah, S.I. Effect of reducing atmosphere on the magnetism of Zn1−xCoxO (0 ≤ x ≤ 0.10) nanoparticles. Nanotechnology 2006, 17, 2675. [Google Scholar] [CrossRef] [PubMed]
  49. Yang, L.W.; Wu, X.L.; Qiu, T. Synthesis and magnetic properties of Zn1−xCoxO nanorods. J. Appl. Phys. 2006, 99, 074303. [Google Scholar] [CrossRef]
  50. Sharrouf, M.; Awad, R.; Marhaba, S.; El-Said Bakeer, D. Structural, Optical and Room Temperature Magnetic Study of Mn-Doped ZnO Nanoparticles. Nano 2016, 11, 1650042. [Google Scholar] [CrossRef]
  51. Fu, J.; Pen, X.; Yan, S.; Gong, Y.; Tan, Y.; Liang, R.; Du, R.; Xing, X.; Mo, G.; Chen, Z.; et al. Synthesis and structural characterization of ZnO doped with Co. J. Alloys Compd. 2013, 558, 212–221. [Google Scholar] [CrossRef]
  52. Omri, K.; El Ghoul, J.; Lemine, O.M.; Bououdina, M.; Zhang, B.; El Mir, L. Magnetic and optical properties of manganese doped ZnO nanoparticles synthesized by sol–gel technique. Superlattices Microstruct. 2013, 60, 139–147. [Google Scholar] [CrossRef]
  53. Sharma, V.K.; Najim, M.; Srivastava, A.K.; Varma, G.D. Structural and magnetic studies on transition metal (Mn, Co) doped ZnO nanoparticles. J. Magn. Magn. Mater. 2012, 324, 683–689. [Google Scholar] [CrossRef]
  54. Blasco, J.; Bartolomé, F.; García, L.M.; García, J.J. Extrinsic origin of ferromagnetism in doped ZnO. J. Mater. Chem. 2006, 16, 2282–2288. [Google Scholar] [CrossRef] [Green Version]
  55. Xu, X.; Cao, C. Hydrothermal synthesis of Co-doped ZnO flakes with room temperature ferromagnetism. J. Alloys Compd. 2010, 501, 265–268. [Google Scholar] [CrossRef]
  56. Mandal, S.K.; Das, A.K.; Natha, T.K. Microstructural and magnetic properties of ZnO:TM (TM = Co, Mn) diluted magnetic semiconducting nanoparticles. J. Appl. Phys. 2006, 100, 104315. [Google Scholar] [CrossRef]
  57. Kuryliszyn-Kudelska, I.; Dobrowolski, W.D.; Kilański, Ł.; Hadžić, B.; Romčević, N.; Sibera, D.; Narkiewicz, U.; Dziawa, P. Magnetic properties of nanocrystalline ZnO doped with MnO and CoO. J. Phys. Conf. Ser. 2010, 200, 072058. [Google Scholar] [CrossRef] [Green Version]
  58. Kuryliszyn-Kudelska, I.; Dobrowolski, W.; Arciszewska, M.; Romčević, N.; Romčević, M.; Hadžić, B.; Sibera, D.; Narkiewicz, U.; Lojkowski, W. Transition metals in ZnO nanocrystals: Magnetic and structural properties. Sci. Sinter. 2013, 45, 31–48. [Google Scholar] [CrossRef]
  59. Kuryliszyn-Kudelska, I.; Dobrowolski, W.; Arciszewska, M.; Romčević, N.; Romčević, M.; Hadžić, B.; Sibera, D.; Narkiewicz, U. Superparamagnetic and ferrimagnetic behavior of nanocrystalline ZnO(MnO). Phys. E Low-Dimens. Syst. Nanostruct. 2018, 98, 10–16. [Google Scholar] [CrossRef]
  60. Żołnierkiewicz, G.; Typek, J.; Guskos, N.; Narkiewicz, U.; Sibera, D. Magnetic resonance study of nanocrystalline 0.10MnO/0.90ZnO. Cent. Eur. J. Phys. 2013, 11, 226–230. [Google Scholar] [CrossRef] [Green Version]
  61. Li, J.H.; Shen, D.Z.; Zhang, J.Y.; Zhao, D.X.; Li, B.S.; Lu, Y.M.; Liu, Y.C.; Fan, X.W. Magnetism origin of Mn-doped ZnO nanoclusters. J. Magn. Magn. Mater. 2006, 302, 118–121. [Google Scholar] [CrossRef]
  62. Samanta, K.; Bhattacharya, P.; Katiyar, R.S.; Iwamoto, W.; Pagliuso, P.G.; Rettori, C. Raman scattering studies in dilute magnetic semiconductor Zn1−xCoxO. Phys. Rev. B 2006, 73, 245213. [Google Scholar] [CrossRef]
  63. Kuryliszyn-Kudelska, I.; Hadžić, B.; Sibera, D.; Romčević, M.; Romčević, N.; Narkiewicz, U.; Łojkowski, W.; Arciszewska, M.; Dobrowolski, W. Magnetic properties of ZnO(Co) nanocrystals. J. Alloys Compd. 2013, 561, 247–251. [Google Scholar] [CrossRef]
  64. Typek, J.; Guskos, N.; Zolnierkiewicz, G.; Sibera, D.; Narkiewicz, U. Magnetic resonance study of Co-doped ZnO nanomaterials: A case of high doping. Rev. Adv. Mater. Sci. 2017, 50, 76–87. [Google Scholar]
  65. Birajdar, S.D.; Alange, R.C.; More, S.D.; Murumkar, V.D.; Jadhav, K.M. Sol-gel Auto Combustion Synthesis, Structural and Magnetic Properties of Mn doped ZnO Nanoparticles. Procedia Manuf. 2018, 20, 174–180. [Google Scholar] [CrossRef]
  66. Luo, X.; Lee, W.T.; Xing, G.; Bao, N.; Yonis, A.; Chu, D.; Lee, J.; Ding, J.; Li, S.; Yi, J. Ferromagnetic ordering in Mn-doped ZnO nanoparticles. Nanoscale Res. Lett. 2014, 9, 625. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Hsu, H.S.; Huang, J.C.A.; Huang, Y.H.; Liao, Y.F.; Lin, M.Z.; Lee, C.H.; Lee, J.F.; Chen, S.F.; Lai, L.Y.; Liu, C.P. Evidence of oxygen vacancy enhanced room-temperature ferromagnetism in Co-doped ZnO. Appl. Phys. Lett. 2006, 88, 242507. [Google Scholar] [CrossRef]
  68. da Silva, R.T.; Mesquita, A.; de Zevallos, A.O.; Chiaramonte, T.; Gratens, X.; Chitta, V.A.; Morbec, J.M.; Rahman, G.; García-Suárez, V.M.; Doriguetto, A.C.; et al. Multifunctional nanostructured Co-doped ZnO: Co spatial distribution and correlated magnetic properties. Phys. Chem. Chem. Phys. 2018, 20, 20257–20269. [Google Scholar] [CrossRef] [PubMed]
  69. Shatnawi, M.; Alsmadi, A.M.; Bsoul, I.; Salameh, B.; Mathai, M.; Alnawashi, G.; Alzoubi, G.M.; Al-Dweri, F.; Bawa’aneh, M.S. Influence of Mn doping on the magnetic and optical properties of ZnO nanocrystalline particles. Results Phys. 2016, 6, 1064–1071. [Google Scholar] [CrossRef]
  70. Mamani, N.C.; da Silva, R.T.; de Zevallos, A.O.; Cotta, A.A.C.; Macedo, W.A.D.; Li, M.S.; Bernardi, M.I.B.; Doriguetto, A.C.; de Carvalho, H.B. On the nature of the room temperature ferromagnetism in nanoparticulate co-doped ZnO thin films prepared by EB-PVD. J. Alloys Compd. 2017, 695, 2682–2688. [Google Scholar] [CrossRef]
  71. Othman, A.A.; Osman, M.A.; Ibrahim, E.M.M.; Ali, M.A.; Abd-Elrahim, A.G. Mn-doped ZnO nanocrystals synthesized by sonochemical method: Structural, photoluminescence, and magnetic properties. Mater. Sci. Eng. B 2017, 219, 1–9. [Google Scholar] [CrossRef]
  72. Martínez, B.; Sandiumenge, F.; Balcells, L.; Arbiol, J.; Sibieude, F.; Monty, C. Structure and magnetic properties of Co-doped ZnO nanoparticles. Phys. Rev. B 2005, 72, 165202. [Google Scholar] [CrossRef]
  73. 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] [PubMed] [Green Version]
  74. Zhang, J.; Tse, K.; Wong, M.; Zhang, Y.; Zhu, J. A brief review of co-doping. Front. Phys. 2016, 11, 117405. [Google Scholar] [CrossRef]
  75. Katayama-Yoshida, H.; Nishimatsu, T.; Yamamoto, T.; Orita, N. Codoping method for the fabrication of low-resistivity wide band-gap semiconductors in p-type GaN, p-type AlN and n-type diamond: Prediction versus experiment. J. Phys. Condens. Matter 2001, 13, 8901–8914. [Google Scholar] [CrossRef]
  76. Eringen, A.C.; Maugin, G.A. Electrodynamics of Continua I: Foundations and Solid Media; Springer: New York, NY, USA, 1990; ISBN 978-1-4612-3226-1. [Google Scholar]
  77. Jung, S.W.; An, S.J.; Yi, G.C.; Jung, C.U.; Lee, S.; Cho, S. Ferromagnetic properties of Zn1−xMnxO epitaxial thin films. Appl. Phys. Lett. 2002, 80, 4561–4563. [Google Scholar] [CrossRef]
  78. Jaćimović, J.; Micković, Z.; Gaál, R.; Smajda, R.; Vâju, C.; Sienkiewicz, A.; Forró, L.; Magrez, A. Synthesis, electrical resistivity, thermo-electric power and magnetization of cubic ZnMnO3. Solid State Commun. 2011, 151, 487–490. [Google Scholar] [CrossRef]
  79. Blasco, J.; García, J. Stable cubic spinels in the Zn–Mn–O system in air. J. Solid State Chem. 2006, 179, 2199–2205. [Google Scholar] [CrossRef]
  80. Han, S.J.; Jang, T.H.; Kim, Y.B.; Park, B.G.; Park, J.H.; Jeong, Y.H. Magnetism in Mn-doped ZnO bulk samples prepared by solid state reaction. Appl. Phys. Lett. 2003, 83, 920–922. [Google Scholar] [CrossRef] [Green Version]
  81. Christensen, A.N.; Ollivier, G. Hydrothermal preparation and low temperature magnetic properties of Mn(OH)2. Solid State Commun. 1972, 10, 609–614. [Google Scholar] [CrossRef]
  82. Norton, D.P.; Overberg, M.E.; Pearton, S.J.; Pruessner, K.; Budai, J.D.; Boatner, L.A.; Chisholm, M.F.; Lee, J.S.; Khim, Z.G.; Park, Y.D.; et al. Ferromagnetism in cobalt-implanted ZnO. Appl. Phys. Lett. 2003, 83, 5488–5490. [Google Scholar] [CrossRef]
  83. Brumage, W.H.; Dorman, C.F.; Quade, C.R. Temperature-dependent paramagnetic susceptibilities of Cu2+ and Co2+ as dilute impurities in ZnO. Phys. Rev. B Condens. Matter 2001, 63, 104411. [Google Scholar] [CrossRef]
  84. Cho, Y.M.; Choo, W.K.; Kim, H.; Kim, D.; Ihm, Y.E. Effects of rapid thermal annealing on the ferromagnetic properties of sputtered Zn1−x(Co0.5Fe0.5)xO thin films. Appl. Phys. Lett. 2002, 80, 3358. [Google Scholar] [CrossRef]
  85. Seidov, Z.; Açıkgöz, M.; Kazan, S.; Mikailzade, F. Magnetic properties of Co3O4 polycrystal powder. Ceram. Int. 2016, 42, 12928–12931. [Google Scholar] [CrossRef]
  86. Singh, V.; Major, D.T. Electronic Structure and Bonding in Co-Based Single and Mixed Valence Oxides: A Quantum Chemical Perspective. Inorg. Chem. 2016, 55, 3307–3315. [Google Scholar] [CrossRef] [PubMed]
  87. Mariappan, C.R.; Kumara, R.; Vijaya Prakash, G. Functional properties of ZnCo2O4 nano-particles obtained by thermal decomposition of a solution of binary metal nitrates. RSC Adv. 2015, 5, 26843–26849. [Google Scholar] [CrossRef]
  88. Gupta, A.; Tiwari, S.D. Magnetic Properties of Undoped and Al doped Layered α-Co(OH)2. Phys. B Condens. Matter 2017, 525, 21–25. [Google Scholar] [CrossRef]
  89. Yan, L.; Ong, C.K.; Rao, X.S. Magnetic order in Co-doped and (Mn, Co) codoped ZnO thin films by pulsed laser deposition. J. Appl. Phys. 2004, 96, 508–511. [Google Scholar] [CrossRef]
  90. Gu, Z.B.; Yuan, C.S.; Lu, M.H.; Wang, J.; Wu, D.; Zhang, S.T.; Zhu, S.N.; Zhu, Y.Y.; Chena, Y.F. Magnetic and transport properties of (Mn, Co)-codoped ZnO films prepared by radio-frequency magnetron cosputtering. J. Appl. Phys. 2005, 98, 053908. [Google Scholar] [CrossRef]
  91. Gu, Z.-B.; Lu, M.-H.; Wang, J.; Du, C.-L.; Yuan, C.-S.; Wu, D.; Zhang, S.-T.; Zhu, Y.-Y.; Zhu, S.-N.; Chen, Y.-F. Optical properties of (Mn, Co) co-doped ZnO films prepared by dual-radio frequency magnetron sputtering. Thin Solid Films 2006, 515, 2361–2365. [Google Scholar] [CrossRef]
  92. Du, C.L.; Gu, Z.B.; You, Y.M.; Kasim, J.; Yu, T.; Shen, Z.X.; Ni, Z.H.; Ma, Y.; Cheng, G.X.; Chen, Y.F. Resonant Raman spectroscopy of (Mn,Co)-codoped ZnO films. J. Appl. Phys. 2008, 103, 023521. [Google Scholar] [CrossRef]
  93. 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]
  94. Nirmala, M.; Smitha, P.; Anukaliani, A. Optical and electrical properties of undoped and (Mn, Co) co-doped ZnO nanoparticles synthesized by DC thermal plasma method. Superlattices Microstruct. 2011, 50, 563–571. [Google Scholar] [CrossRef]
  95. Nirmala, M.; Anukaliani, A. Synthesis and characterization of undoped and TM (Co, Mn) doped ZnO nanoparticles. Mater. Lett. 2011, 65, 2645–2648. [Google Scholar] [CrossRef]
  96. Li, H.; Huang, Y.; Zhang, Q.; Qiao, Y.; Gu, Y.; Liu, J.; Zhang, Y. Facile synthesis of highly uniform Mn/Co-codoped ZnO nanowires: Optical, electrical, and magnetic properties. Nanoscale 2011, 3, 654–660. [Google Scholar] [CrossRef] [PubMed]
  97. Inamdar, D.Y.; Lad, A.D.; Pathak, A.K.; Dubenko, I.; Ali, N.; Mahamuni, S. Ferromagnetism in ZnO Nanocrystals: Doping and Surface Chemistry. J. Phys. Chem. C 2010, 114, 1451–1459. [Google Scholar] [CrossRef]
  98. Phan, L.; Vincent, R.; Cherns, D.; Dan, N.H.; Yu, S.C. Characterization of (Mn, Co)-Codoped ZnO Nanorods Prepared by Thermal Diffusion. IEEE Trans. Magn. 2009, 45, 2435–2438. [Google Scholar] [CrossRef]
  99. Naeem, M.; Hasanain, S.K. Role of donor defects in stabilizing room temperature ferromagnetism in (Mn, Co) co-doped ZnO nanoparticles. J. Phys. Condens. Matter 2012, 24, 245305. [Google Scholar] [CrossRef] [PubMed]
  100. Yin-Hua, Y.; Xi, C.Q. Infrared emissivities of Mn, Co co-doped ZnO powders. Chin. Phys. B 2012, 21, 124205. [Google Scholar] [CrossRef]
  101. Sato, W.; Kano, Y.; Suzuki, T.; Nakagawa, M.; Kobayashi, Y. Local fields in Co and Mn Co-doped ZnO. Hyperfine Interact. 2016, 237, 28. [Google Scholar] [CrossRef]
  102. Li, H.; Liu, H.; Zheng, Z.J. Magnetic behavior of Co–Mn co-doped ZnO nanoparticles. Magn. Magn. Mater. 2014, 372, 37–40. [Google Scholar] [CrossRef]
  103. Neena, D.; Shah, A.H.; Deshmukh, K.; Ahmad, H.; Fu, D.J.; Kondamareddy, K.K.; Kumar, P.; Dwivedi, R.K.; Sing, V. Influence of (Co-Mn) co-doping on the microstructures, optical properties of sol-gel derived ZnO nanoparticles. Eur. Phys. J. D 2016, 70, 53. [Google Scholar] [CrossRef]
  104. Sivaselvan, S.; Muthukumaran, S.; Ashokkumar, M. Influence of Co-doping on the structural, optical and morphological properties of Zn0.96Mn0.04O nanoparticles by sol–gel method. Opt. Mater. 2014, 36, 797–803. [Google Scholar] [CrossRef]
  105. Abdullahi, S.S.; Köseog, Y.; Güner, S.; Kazan, S.; Kocaman, B.; Ndikilar, C.E. Synthesis and characterization of Mn and Co codoped ZnO nanoparticles. Superlattices Microstruct. 2015, 83, 342–352. [Google Scholar] [CrossRef]
  106. Köseoğlu, Y. Fuel aided rapid synthesis and room temperature ferromagnetism of M0.1Co0.1Zn0.8O (M = Mn, Ni, Fe and Cu) DMS nanoparticles. Ceram. Int. 2016, 42, 9190–9195. [Google Scholar] [CrossRef]
  107. Pazhanivelu, V.; Blessington Selvadurai, A.P.; Zhao, Y.; Thiyagarajan, R.; Murugaraj, R. Room temperature ferromagnetism in Mn doped ZnO: Co nanoparticles by co-precipitation method. Phys. B Condens. Matter 2016, 481, 91–96. [Google Scholar] [CrossRef]
  108. Pazhanivelu, V.; Blessington Selvadurai, A.P.; Murugaraj, R. Sintering Effect on Structural, Optical and Unusual Magnetic Behaviour in Zn0.95Co0.05O-Based DMS Materials. J. Supercond. Nov. Magn. 2015, 28, 2575–2581. [Google Scholar] [CrossRef]
  109. Pazhanivelu, V.; Blessington Selvadurai, A.P.; Murugaraj, R. Room temperature magnetic behaviour of Mn codoping in ZnO: Co nanoparticles synthesized by co-precipitation method. J. Mater. Sci. Mater. Electron. 2018, 29, 3087–3094. [Google Scholar] [CrossRef]
  110. Yakout, S.M.; El-Sayed, A.M. Synthesis, Structure, and Room Temperature Ferromagnetism of Mn and/or Co Doped ZnO Nanocrystalline. J. Supercond. Nov. Magn. 2016, 29, 1593–1599. [Google Scholar] [CrossRef] [Green Version]
  111. Karthika, K.; Ravichandran, K. Enhancing the magnetic and antibacterial properties of ZnO nanopowders through Mn+Co doping. Ceram. Int. 2015, 41, 7944–7951. [Google Scholar] [CrossRef]
  112. 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]
  113. Khan, R.; de Araujo, C.I.L.; Khan, T.; Khan, A.; Ullah, B.; Fashu, S. Influence of oxygen vacancies on the structural, dielectric, and magnetic properties of (Mn, Co) co-doped ZnO nanostructures. J. Mater. Sci. Mater. Electron. 2018, 29, 9785–9795. [Google Scholar] [CrossRef]
  114. Khan, R.; Fashu, Z.S.; Rehman, Y.U.; Khan, A.; Rahman, M.U. Structure and magnetic properties of (Co, Mn) co-doped ZnO diluted magnetic semiconductor nanoparticles. J. Mater. Sci. Mater. Electron. 2018, 29, 32–37. [Google Scholar] [CrossRef]
  115. Li, G.R.; Qu, D.L.; Zhao, W.X.; Tong, Y.X. Electrochemical deposition of (Mn, Co)-codoped ZnO nanorod arrays without any template. Electrochem. Commun. 2007, 9, 1661–1666. [Google Scholar] [CrossRef]
  116. Dai, J.; Meng, C.; Li, Q. First-principles study on the magnetism of Mn and Co codoped ZnO. Phys. B Condens. Matter 2013, 409, 5–9. [Google Scholar] [CrossRef]
  117. Stashans, A.; Rivera, K. Electronic and Magnetic Properties of Co- and Mn-codoped ZnO by Density Functional Theory. Chin. Phys. Lett. 2016, 33, 097102. [Google Scholar] [CrossRef]
  118. Liu, Z.; Yuan, X.; Yang, P. Investigation on electronic and magnetic properties of Co and Mn in ZnO with different doping types. J. Magn. Magn. Mater. 2018, 461, 1–5. [Google Scholar] [CrossRef]
  119. Horikoshi, S.; Serpone, N. Microwaves in Nanoparticle Synthesis: Fundamentals and Applications; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2013; ISBN 9783527331970. [Google Scholar]
  120. Chikan, V.; McLaurin, E.J. Rapid Nanoparticle Synthesis by Magnetic and Microwave Heating. Nanomaterials 2016, 6, 85. [Google Scholar] [CrossRef] [PubMed]
  121. Saleh, T.A.; Majeed, S.; Nayak, A.; Bhushan, B. Principles and Advantages of Microwave—Assisted Methods for the Synthesis of Nanomaterials for Water Purification. In Advanced Nanomaterials for Water Engineering, Treatment, and Hydraulics, 2nd ed.; Saleh, T.A., Ed.; IGI Global: Hershey, PA, USA, 2017; pp. 40–57. ISBN -10 1522521364. [Google Scholar]
  122. Motshekga, S.C.; Pillai, S.K.; Ray, S.S.; Jalama, K.; Krause, R.W.M. Recent Trends in the Microwave-Assisted Synthesis of Metal Oxide Nanoparticles Supported on Carbon Nanotubes and Their Applications. J. Nanomater. 2012, 2012, 691503. [Google Scholar] [CrossRef]
  123. Bilecka, I.; Niederberger, M. Microwave chemistry for inorganic nanomaterials synthesis. Nanoscale 2010, 2, 1358–1374. [Google Scholar] [CrossRef] [PubMed]
  124. Gawande, M.B.; Shelke, S.N.; Zboril, R.; Varma, R.S. Microwave-Assisted Chemistry: Synthetic Applications for Rapid Assembly of Nanomaterials and Organics. Acc. Chem. Res. 2014, 47, 1338–1348. [Google Scholar] [CrossRef] [PubMed]
  125. Koltsov, I.; Prześniak-Welenc, M.; Wojnarowicz, J.; Rogowska, A.; Mizeracki, J.; Malysa, M.; Kimmel, G. Thermal and physical properties of ZrO2-AlO(OH) nanopowders synthesised by microwave hydrothermal method. J. Therm. Anal. Calorim. 2017, 13, 2273–2284. [Google Scholar] [CrossRef]
  126. Guenin, E. Microwave Engineering of Nanomaterials: From Mesoscale to Nanoscale, 1st ed.; Pan Stanford Publishing Pte. Ltd.: Boca Raton, FL, USA, 2016; ISBN 9789814669429. [Google Scholar]
  127. Polshettiwar, V.; Nadagouda, M.N.; Varma, R.S. Microwaveassisted chemistry: A rapid and sustainable route to synthesis of organics and nanomaterials. Aust. J. Chem. 2009, 62, 16–26. [Google Scholar] [CrossRef]
  128. Wojnarowicz, J.; Chudoba, T.; Majcher, A.; Łojkowski, W. Microwaves Applied to Hydrothermal Synthesis of Nanoparticles. In Microwave Chemistry, 1st ed.; Cravotto, G., Carnaroglio, D., Eds.; De Gruyter: Berlin, Germany; Boston, MA, USA, 2017; pp. 205–224. ISBN 9783110479935. [Google Scholar]
  129. Lojkowski, W.; Leonelli, C.; Chudoba, T.; Wojnarowicz, J.; Majcher, A.; Mazurkiewicz, A. High-Energy-low-temperature technologies for the synthesis of nanoparticles: Microwaves and high pressure. Inorganics 2014, 2, 606–619. [Google Scholar] [CrossRef]
  130. Kołodziejczak-Radzimska, A.; Jesionowski, T. Zinc Oxide-From Synthesis to Application: A Review. Materials 2014, 7, 2833–2881. [Google Scholar] [CrossRef] [PubMed]
  131. Hasanpoor, M.; Aliofkhazraei, M.; Delavari, H. Microwave-assisted Synthesis of Zinc Oxide Nanoparticles. Proc. Mater. Sci. 2015, 11, 320–325. [Google Scholar] [CrossRef]
  132. Shao, D.; Wei, Q. Microwave-Assisted Rapid Preparation of Nano-ZnO/Ag Composite Functionalized Polyester Nonwoven Membrane for Improving Its UV Shielding and Antibacterial Properties. Materials 2018, 11, 1412. [Google Scholar] [CrossRef] [PubMed]
  133. Rana, A.U.; Kang, M.; Kim, H.S. Microwave-assisted Facile and Ultrafast Growth of ZnO Nanostructures and Proposition of Alternative Microwave-assisted Methods to Address Growth Stoppage. Sci. Rep. 2016, 6, 24870. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Rana, A.H.S.; Chang, S.B.; Kim, H.S. NH4OH-Oriented and pH-Dependent Growth of ZnO Nanostructures via Microwave-Assisted Growth Method. J. Nanosci. Nanotechnol. 2018, 18, 2125–2127. [Google Scholar] [CrossRef] [PubMed]
  135. Pimentel, A.; Ferreira, S.H.; Nunes, D.; Calmeiro, T.; Martins, R.; Fortunato, E. Microwave Synthesized ZnO Nanorod Arrays for UV Sensors: A Seed Layer Annealing Temperature Study. Materials 2016, 9, 299. [Google Scholar] [CrossRef] [PubMed]
  136. Pimentel, A.; Nunes, D.; Duarte, P.; Rodrigues, J.; Costa, F.M.; Monteiro, T.; Martins, R.; Fortunato, E. Synthesis of Long ZnO Nanorods under Microwave Irradiation or Conventional Heating. J. Phys. Chem. C 2014, 118, 14629–14639. [Google Scholar] [CrossRef]
  137. Barreto, G.P.; Morales, G.; Quintanilla Ma, L.L. Microwave Assisted Synthesis of ZnO Nanoparticles: Effect of Precursor Reagents, Temperature, Irradiation Time, and Additives on Nano-ZnO Morphology Development. J. Mater. 2013, 2013, 478681. [Google Scholar] [CrossRef]
  138. Barreto, G.; Morales, G.; Cañizo, A.; Eyler, N. Microwave Assisted Synthesis of ZnO Tridimensional Nanostructures. Proc. Mater. Sci. 2015, 8, 535–540. [Google Scholar] [CrossRef]
  139. Ghosh, S.; Chakraborty, J. Rapid synthesis of zinc oxide nanoforest: Use of microwave and forced seeding. Mater. Res. Express 2016, 3, 125004. [Google Scholar] [CrossRef]
  140. Wojnarowicz, J.; Opalinska, A.; Chudoba, T.; Gierlotka, S.; Mukhovskyi, R.; Pietrzykowska, E.; Sobczak, K.; Lojkowski, W. Effect of water content in ethylene glycol solvent on the size of ZnO nanoparticles prepared using microwave solvothermal synthesis. J. Nanomater. 2016, 2016, 2789871. [Google Scholar] [CrossRef]
  141. Wojnarowicz, J.; Chudoba, T.; Koltsov, I.; Gierlotka, S.; Dworakowska, S.; Lojkowski, W. Size control mechanism of ZnO nanoparticles obtained in microwave solvothermal synthesis. Nanotechnology 2018, 29, 065601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  142. Mohammadi, E.; Aliofkhazraei, M.; Hasanpoor, M.; Chipara, M. Hierarchical and Complex ZnO Nanostructures by Microwave-Assisted Synthesis: Morphologies, Growth Mechanism and Classification. Crit. Rev. Solid State Mater. Sci. 2018, 43, 475–541. [Google Scholar] [CrossRef]
  143. Majcher, A.; Wiejak, J.; Przybylski, J.; Chudoba, T.; Wojnarowicz, J. A novel reactor for microwave hydrothermal scale-up nanopowder synthesis. Int. J. Chem. React. Eng. 2013, 11, 361–368. [Google Scholar] [CrossRef]
  144. Cravotto, G.; Carnaroglio, D. Microwave Chemistry, 1st ed.; De Gruyter: Berlin, Germany; Boston, MA, USA, 2017; ISBN 9783110479935. [Google Scholar]
  145. Rinaldi, L.; Carnaroglio, D.; Rotolo, L.; Cravotto, G. A Microwave-Based Chemical Factory in the Lab: From Milligram to Multigram Preparations. J. Chem. 2015, 2015, 879531. [Google Scholar] [CrossRef]
  146. Dąbrowska, S.; Chudoba, T.; Wojnarowicz, J.; Łojkowski, W. Current Trends in the Development of Microwave Reactors for the Synthesis of Nanomaterials in Laboratories and Industries: A Review. Crystals 2018, 8, 379. [Google Scholar] [CrossRef]
  147. Dąbrowska, S.; Chudoba, T.; Wojnarowicz, J.; Łojkowski, W. Problems of exploitations of microwave reactors for nanoparticles synthesis. J. Mach. Construct. Maint. 2018, 110, 41–47. [Google Scholar]
  148. Wojnarowicz, J.; Chudoba, T.; Gierlotka, S.; Lojkowski, W. Effect of Microwave Radiation Power on the Size of Aggregates of ZnO NPs Prepared Using Microwave Solvothermal Synthesis. Nanomaterials 2018, 8, 343. [Google Scholar] [CrossRef] [PubMed]
  149. Wojnarowicz, J.; Mukhovskyi, R.; Pietrzykowska, E.; Kusnieruk, S.; Mizeracki, J.; Lojkowski, W. Microwave solvothermal synthesis and characterization of manganese-doped ZnO nanoparticles. Beilstein J. Nanotechnol. 2016, 7, 721–732. [Google Scholar] [CrossRef] [PubMed]
  150. Wojnarowicz, J.; Kuśnieruk, S.; Chudoba, T.; Mizeracki, J.; Łojkowski, W. Microwave solvothermal synthesis of Co-doped ZnO nanoparticles. Glass Ceram. 2015, 3, 8–13. [Google Scholar]
  151. Wojnarowicz, J.; Chudoba, T.; Gierlotka, S.; Sobczak, K.; Lojkowski, W. Size Control of Cobalt-Doped ZnO Nanoparticles Obtained in Microwave Solvothermal Synthesis. Crystals 2018, 8, 179. [Google Scholar] [CrossRef]
  152. Lojkowski, W.; Gedanken, A.; Grzanka, E.; Opalinska, A.; Strachowski, T.; Pielaszek, R.; Tomaszewska-Grzeda, A.; Yatsunenko, S.; Godlewski, M.; Matysiak, H.; et al. Solvothermal synthesis of nanocrystalline zinc oxide doped with Mn2+, Ni2+, Co2+ and Cr3+ ions. J. Nanopart. Res. 2009, 11, 1991–2002. [Google Scholar] [CrossRef]
  153. Pielaszek, R. FW15/45M method for determination of the grain size distribution from powder diffraction line profile. J. Alloys Compd. 2004, 37, 128–132. [Google Scholar] [CrossRef]
  154. Nanopowder XRD Processor Demo, pre⋅α⋅ver.0.0.8, © Pielaszek Research. Available online: http://science24.com/xrd/ (accessed on 10 January 2018).
  155. FW1/5 4/5M Method of Evaluation of Grain Size Distribution by Powder Diffraction. Available online: http://science24.com/fw145m/ (accessed on 10 January 2018).
  156. Zhang, Q.; Park, K.; Cao, G. Synthesis of ZnO Aggregates and Their Application in Dye-sensitized Solar Cells. Mater. Matters 2014, 5, 32–39. [Google Scholar]
  157. Poul, L.; Jouini, N.; Fiévet, F. Layered Hydroxide Metal Acetates (Metal = Zinc, Cobalt, and Nickel): Elaboration via Hydrolysis in Polyol Medium and Comparative Study. Chem. Mater. 2000, 12, 3123–3132. [Google Scholar] [CrossRef]
  158. O’Keeffe, M.; Hyde, B.G. Crystal Structures: I. Patterns and Symmetry, 1st ed.; Mineralogical Society of America: Washington, DC, USA, 1996; ISBN 0939950405. [Google Scholar]
  159. Rao, M.S.R.; Okada, T. ZnO Nanocrystals and Allied Materials; Springer: New Delhi, India, 2014; ISBN 978-81-322-1159-4. [Google Scholar]
  160. Kück, S.; Werheit, H. Non-Tetrahedrally Bonded Binary Compounds II; Springer: Berlin/Heidelberg, Germany, 2000; ISBN 978-3-540-64966-3. [Google Scholar]
  161. Sarbas, B.; Töpper, W. Mn Manganese; Springer: Berlin/Heidelberg, Germany, 1993; ISBN 978-3-662-08907-1. [Google Scholar]
  162. Saravanan, R. Solid Oxide Fuel Cell (SOFC) Materials; Materials Research Forum LLC: Millersville, PA, USA, 2018; ISBN 978-1-945291-50-0. [Google Scholar]
  163. Olesik, J.W. Elemental analysis using ICP-OES and ICP/MS. Anal. Chem. 1991, 63, 12A–21A. [Google Scholar] [CrossRef]
  164. Barsoum, M.W. Fundamentals of Ceramics, 1st ed.; Taylor & Francis Group: New York, NY, USA; Northampton, UK, 2002; ISBN 9780750309028. [Google Scholar]
  165. Opalinska, A.; Malka, I.; Dzwolak, W.; Chudoba, T.; Presz, A.; Lojkowski, W. Size-dependent density of zirconia nanoparticles. Beilstein J. Nanotechnol. 2015, 6, 27–35. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  166. Kusnieruk, S.; Wojnarowicz, S.; Chodara, A.; Chudoba, T.; Gierlotka, S.; Lojkowski, W. Influence of hydrothermal synthesis parameters on the properties of hydroxyapatite nanoparticles. Beilstein J. Nanotechnol. 2016, 7, 1586–1601. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  167. Wejrzanowski, T.; Pielaszek, R.; Opalińska, A.; Matysiak, H.; Łojkowski, W.; Kurzydłowski, K.J. Quantitative methods for nanopowders characterization. Appl. Surf. Sci. 2006, 253, 204–208. [Google Scholar] [CrossRef]
  168. Zając, M.; Gosk, J.; Grzanka, E.; Kamińska, M.; Twardowski, A.; Strojek, B.; Szyszko, T.; Podsiadło, S. Possible origin of ferromagnetism in (Ga,Mn)N. J. Appl. Phys. 2003, 93, 4715–4717. [Google Scholar] [CrossRef]
  169. Zając, M.; Gosk, J.; Grzanka, E.; Stelmakh, S.; Palczewska, M.; Wysmołek, A.; Korona, K.; Kamińska, M.; Twardowski, A. Ammonothermal synthesis of GaN doped with transition metal ions (Mn, Fe, Cr). J. Alloys Compd. 2008, 456, 324–338. [Google Scholar] [CrossRef]
  170. Lawniczak-Jablonska, K.; Wolska, A.; Klepka, M.T.; Kret, S.; Gosk, J.; Twardowski, A.; Wasik, D.; Kwiatkowski, A.; Kurowska, B.; Kowalski, B.J.; et al. Magnetic properties of MnSb inclusions formed in GaSb matrix directly during molecular beam epitaxial growth. J. Appl. Phys. 2011, 109, 074308. [Google Scholar] [CrossRef]
  171. Gaj, J.A.; Planel, R.; Fishman, G. Relation of magneto-optical properties of free excitons to spin alignment of Mn2+ ions in Cd1−xMnxTe. Solid State Commun. 1979, 29, 435–438. [Google Scholar] [CrossRef]
  172. Gaj, J.A. On the physical meaning of the modified Brillouin function. Acta Phys. Pol. 1988, 73, 463–466. [Google Scholar]
Figure 1. Diagram showing the course of the microwave synthesis of codoped ZnO NPs and undoped ZnO NPs on the example of Zn0.98Mn0.01Co0.01O sample. Experimental data obtained in the microwave reactor Model 02-02.
Figure 1. Diagram showing the course of the microwave synthesis of codoped ZnO NPs and undoped ZnO NPs on the example of Zn0.98Mn0.01Co0.01O sample. Experimental data obtained in the microwave reactor Model 02-02.
Crystals 08 00410 g001
Figure 2. SEM images of NPs: (a,b) ZnO; (c,d) Zn0.98Co0.01Mn0.01O; (e,f) Zn0.90Co0.05Mn0.05O; (g,h) Zn0.8Co0.1Mn0.1O.
Figure 2. SEM images of NPs: (a,b) ZnO; (c,d) Zn0.98Co0.01Mn0.01O; (e,f) Zn0.90Co0.05Mn0.05O; (g,h) Zn0.8Co0.1Mn0.1O.
Crystals 08 00410 g002
Figure 3. SEM images of NPs: (ad) Zn0.7Co0.15Mn0.15O.
Figure 3. SEM images of NPs: (ad) Zn0.7Co0.15Mn0.15O.
Crystals 08 00410 g003
Figure 4. XRD diffraction patterns of Zn(1−x−y)MnxCoyO NPs (where x = y), with the nominal content of dopants in the solution being 0, 1, 5, 10, 15 mol %.
Figure 4. XRD diffraction patterns of Zn(1−x−y)MnxCoyO NPs (where x = y), with the nominal content of dopants in the solution being 0, 1, 5, 10, 15 mol %.
Crystals 08 00410 g004
Figure 5. XRD diffraction pattern of the Zn0.7Co0.15Mn0.15O sample, and its comparison with the standard pattern of Zn5 (OH)7.9(CH3COO)2.1·1.55H2O (JCPDS No. 56-0569) and ZnO (JCPDS No. 36-1451).
Figure 5. XRD diffraction pattern of the Zn0.7Co0.15Mn0.15O sample, and its comparison with the standard pattern of Zn5 (OH)7.9(CH3COO)2.1·1.55H2O (JCPDS No. 56-0569) and ZnO (JCPDS No. 36-1451).
Crystals 08 00410 g005
Figure 6. Lattice parameters versus nominal dopants content of Zn(1−x−y)MnxCoyO NP samples.
Figure 6. Lattice parameters versus nominal dopants content of Zn(1−x−y)MnxCoyO NP samples.
Crystals 08 00410 g006
Figure 7. Photos of NP suspensions in ethylene glycol immediately after microwave solvothermal synthesis: (a) Zn(1−x−y)MnxCoyO, (b) Zn(1−y)CoyO [150], (c) Zn(1−x)MnxO [149].
Figure 7. Photos of NP suspensions in ethylene glycol immediately after microwave solvothermal synthesis: (a) Zn(1−x−y)MnxCoyO, (b) Zn(1−y)CoyO [150], (c) Zn(1−x)MnxO [149].
Crystals 08 00410 g007
Figure 8. Photos of dry powders of Zn(1−x−y)MnxCoyO.
Figure 8. Photos of dry powders of Zn(1−x−y)MnxCoyO.
Crystals 08 00410 g008
Figure 9. Crystallite size distribution of Zn(1−x−y)MnxCoyO NPs: (a) ZnO, (b) Zn0.98Mn0.01Co0.01O, (c) Zn0.90Mn0.05Co0.05O, (d) Zn0.80Mn0.10Co0.10O, and (e) Zn0.70Mn0.15Co0.15O. Data obtained using Nanopowder XRD Processor Demo, pre·α·ver.0.0.8, © Pielaszek Research, http://science24.com/xrd/.
Figure 9. Crystallite size distribution of Zn(1−x−y)MnxCoyO NPs: (a) ZnO, (b) Zn0.98Mn0.01Co0.01O, (c) Zn0.90Mn0.05Co0.05O, (d) Zn0.80Mn0.10Co0.10O, and (e) Zn0.70Mn0.15Co0.15O. Data obtained using Nanopowder XRD Processor Demo, pre·α·ver.0.0.8, © Pielaszek Research, http://science24.com/xrd/.
Crystals 08 00410 g009
Figure 10. Measured magnetisation as a function of the magnetic field: (a) Zn(1−x−y)MnxCoyO—temperature—2 K, (b) Zn(1−x−y)MnxCoyO—temperature 300 K. Red full points—Zn0.70Mn0.15Co0.15O, blue—Zn0.80Mn0.10Co0.10O, green—Zn0.90Mn0.05Co0.05O, black—Zn0.98Mn0.01Co0.01O, brown crosses—pure ZnO.
Figure 10. Measured magnetisation as a function of the magnetic field: (a) Zn(1−x−y)MnxCoyO—temperature—2 K, (b) Zn(1−x−y)MnxCoyO—temperature 300 K. Red full points—Zn0.70Mn0.15Co0.15O, blue—Zn0.80Mn0.10Co0.10O, green—Zn0.90Mn0.05Co0.05O, black—Zn0.98Mn0.01Co0.01O, brown crosses—pure ZnO.
Crystals 08 00410 g010
Figure 11. Magnetisation difference Mexp(B, T = 2 K)—Mexp(B, T = 300 K) as a function of the magnetic field for: (a) Zn0.99Mn0.01O, solid line: fit with Equations (5) and (6); T0 was set to 0 K; (b) Zn0.99Mn0.01O and Zn0.85Mn0.15O (red), solid lines: fits with Equations (5) and (6), with S = 5/2 and T0 as an adjustable parameter (T0 = −0.17 K for Zn0.99Mn0.01O and T0 = −0.29 K for Zn0.85Mn0.15O sample), Zn0.99Co0.01O and Zn0.85Co0.15O (blue), solid lines: fits with Equations (5) and (6), with S = 3/2 and T0 as an adjustable parameter (T0 = −0.24 K for Zn0.99Co0.01O and T0 = −0.95 K for Zn0.85Co0.15O sample).
Figure 11. Magnetisation difference Mexp(B, T = 2 K)—Mexp(B, T = 300 K) as a function of the magnetic field for: (a) Zn0.99Mn0.01O, solid line: fit with Equations (5) and (6); T0 was set to 0 K; (b) Zn0.99Mn0.01O and Zn0.85Mn0.15O (red), solid lines: fits with Equations (5) and (6), with S = 5/2 and T0 as an adjustable parameter (T0 = −0.17 K for Zn0.99Mn0.01O and T0 = −0.29 K for Zn0.85Mn0.15O sample), Zn0.99Co0.01O and Zn0.85Co0.15O (blue), solid lines: fits with Equations (5) and (6), with S = 3/2 and T0 as an adjustable parameter (T0 = −0.24 K for Zn0.99Co0.01O and T0 = −0.95 K for Zn0.85Co0.15O sample).
Crystals 08 00410 g011
Figure 12. Magnetisation Mexp(B, T = 2 K)—Mexp(B, T = 300 K) of codoped ZnO NPs with Mn2+ and Co2+ versus magnetic field: (a) Zn0.70Mn0.15Co0.15O empty points—Mexp(B, T = 2 K), crosses—Mexp(B, T = 300 K), full points Mexp(B, T = 2 K)—Mexp(B, T = 300 K), solid line—fit with Equation (7); (b) Mexp(B, T = 2 K)—Mexp(B, T = 300 K) for Zn0.70Mn0.15Co0.15O (red), Zn0.80Mn0.10Co0.10O (blue), Zn0.90Mn0.05Co0.05O (green), Zn0.98Mn0.01Co0.01O (black). Solid lines—fits with Equation (7), where ratio AMn/ACo was kept as resulting from Table 5. The resulting T0 are the following: T0 = −1.78 K for Zn0.70Mn0.15Co0.15O (red), T0 = −0.42 K for Zn0.80Mn0.10Co0.10O (blue), T0 = −0.37 K for Zn0.90Mn0.05Co0.05O (green), T0 = −0.35 K for Zn0.98Mn0.01Co0.01O (black).
Figure 12. Magnetisation Mexp(B, T = 2 K)—Mexp(B, T = 300 K) of codoped ZnO NPs with Mn2+ and Co2+ versus magnetic field: (a) Zn0.70Mn0.15Co0.15O empty points—Mexp(B, T = 2 K), crosses—Mexp(B, T = 300 K), full points Mexp(B, T = 2 K)—Mexp(B, T = 300 K), solid line—fit with Equation (7); (b) Mexp(B, T = 2 K)—Mexp(B, T = 300 K) for Zn0.70Mn0.15Co0.15O (red), Zn0.80Mn0.10Co0.10O (blue), Zn0.90Mn0.05Co0.05O (green), Zn0.98Mn0.01Co0.01O (black). Solid lines—fits with Equation (7), where ratio AMn/ACo was kept as resulting from Table 5. The resulting T0 are the following: T0 = −1.78 K for Zn0.70Mn0.15Co0.15O (red), T0 = −0.42 K for Zn0.80Mn0.10Co0.10O (blue), T0 = −0.37 K for Zn0.90Mn0.05Co0.05O (green), T0 = −0.35 K for Zn0.98Mn0.01Co0.01O (black).
Crystals 08 00410 g012
Table 1. Magnetic properties of the secondary phase observed in Mn2+-doped ZnO and Co2+-doped ZnO.
Table 1. Magnetic properties of the secondary phase observed in Mn2+-doped ZnO and Co2+-doped ZnO.
PhaseMagnetic PropertiesReference
MnAntiferromagnetic[76]
MnOAntiferromagnetic[77]
MnO2Antiferromagnetic[77]
Mn2O3Antiferromagnetic[77]
Mn3O4Ferromagnetic[77]
ZnMnO3Paramagnetic[78,79]
ZnMn2O4Ferromagnetic[80]
Mn(OH)2Antiferromagnetic[81]
CoFerromagnetic[82]
Co2+clusterFerromagnetic[83]
CoOAntiferromagnetic[84]
Co3O4Antiferromagnetic[85]
Co2O3Antiferromagnetic[86]
ZnCo2O4Ferromagnetic[87]
Co(OH)2Paramagnetic[88]
Table 2. Summary of methods of obtaining Mn2+-and Co2+-codoped ZnO.
Table 2. Summary of methods of obtaining Mn2+-and Co2+-codoped ZnO.
MethodSubstratesConditions during PreparationMorphology;
Magnetic Properties;
Maximum Content of Dopants
Reference
pulsed laser depositionZnO, Co3O4 with MnO2from 400 °C to 600 °C and a low oxygen pressure (5 × 10−5 Pa).films;
room-temperature ferromagnetism;
Zn0.7Mn0.15Co0.15O
[89]
radio-frequency
magnetron co-sputtering and annealing
Zn0.95Mn0.05O, Zn0.80Co0.20O550 °C in argon ambient;
400 °C, 600 °C, 800 °C and 900 °C for 1 h in air
films;
ferromagnetic and paramagnetic behaviour;
Zn0.90Mn0.03Co0.07O
[90,91,92]
autocombustionZn(NO3)2·6H2O, Co(NO3)2·6H2O, Mn(NO3)2, NH2CH2COOH and H2O-nanoparticles;
ferromagnetic;
Zn0.96Mn0.02Co0.02O
[93]
thermal plasma and annealingzinc, cobalt and manganese metal powders450 °C, 550 °C and 650 °C for 1 h, atmospheric airParticles;[94,95]
thermal evaporationZn powders, MnCl2·4H2O, CoCl2·6H2O500 °C with a constant argon flow of 50 sccmnanowires (50–200 nm and lengths up to several tens of microns)[96]
thermal decompositionZn(Ac)2·2H2O, Mn(Ac)2·4H2O, Co(Ac)2·4H2O and C16H33NH2 and C4H9PO(OH)2at 300 °C under a constant nitrogen flownanocrystals (7 nm);
room temperature ferromagnetism
[97]
thermal diffusionZn(NO3)2·6H2O, C6H12N4, Mn and Co (metal powders)at 850 °C in a vacuum at about 1 × 10−3 Torr.nanorods (1–2 µm length and 80–200 nm diameter)[98]
chemical route and annealingZn(Ac)2·2H2O, Mn(Ac)2·4H2O, Co(Ac)2·4H2O and ethylene glycol200 °C for 3 h;
600 °C in (Ar95% H5%) and in air.
nanoparticles (25 nm);
paramagnetic and ferromagnetic behaviour; Zn0.94Mn0.04Co0.02O,
Zn0.94Mn0.02Co0.04O
[99]
solid state reaction and calcinationZnO was mixed with MnCO3 and Co2O (the mixtures were ball milled)2 h in 1000 °C, 1050 °C, 1100 °C, 1150 °C, and 1200 °C in atmospheric airparticles;
Zn0.94Mn0.01Co0.05O
[100]
solid state reaction and calcinationZnO was mixed with Mn and Co (metal powders) with the assistance citric acid1100 °C for 7 day in airroom temperature ferromagnetism;
Zn0.90Mn0.05Co0.05O
[101]
sol–gelZn(Ac)2·2H2O, Mn(Ac)2·4H2O, Co(Ac)2·4H2O ethanol solution as a solvent, and DEA as a stabilizing agent500 °C for 1 h in O2 atmosphere,particles;
exhibited ferromagnetic character;
Zn0.94Mn0.04Co0.02O
[102]
Zn(Ac)2·2H2O, Mn(Ac)2·4H2O, Co(Ac)2·4H2O, NaOH, water and N,N-dimethylformamide,500 °C for 1 h in atmospheric airnanoparticles; ferromagnetism behaviour;
Zn0.7Co0.15Mn0.15O
[103]
Zn(Ac)2·2H2O, Mn(Ac)2·4H2O, Co(Ac)2·4H2O, N,N dimethyl-formamide (DMF)500 °C for 4 h in atmospheric airnanoparticles (20–30 nm);
Zn0.88Mn0.04Co0.08O
[104]
microwave assisted combustion synthesisMn(NO3)2·4H2O, Co(NO3)6·H2O, Zn(NO3)2·6H2O, water, using urea as a fuel800 W for 15 min (kitchen type microwave)nanoparticles (24 nm);
paramagnetic and ferromagnetic behaviour;
Zn0.70Mn0.20Co0.10O
[105]
Mn(NO3)2·4H2O, Co(NO3)6·H2O, Zn(NO3)2·6H2O, water, using urea as a fuel1000 W for 20 min (kitchen type microwave)nanoparticles (31 nm);
room temperature ferromagnetism;
Zn0.80Mn0.10Co0.10O
[106]
co-precipitation and calcinationZn(Ac)2·2H2O, Mn(Ac)2·4H2O, Co(Ac)2·4H2O, (COOH)2·2H2O350 °C, 500 °C and 650 °C in atmospheric airnanoparticles;
ferromagnetism
behaviour
Zn0.90Mn0.05Co0.05O
[107,108,109]
Zn(Ac)2·2H2O, MnCl4·4H2O, CoCl2·6H2O, NH4OH400 °C for 3 h in atmospheric airnanoparticles (28 nm);
ferromagnetic behaviour;
Zn0.98Mn0.01Co0.01O
[110]
Zn(Ac)2·2H2O, Mn(Ac)2·4H2O, Co(Ac)2·4H2O, Na(OH), water550 °C for 3 h in atmospheric airnanoparticles (20 nm); ferromagnetic behaviour;
Zn0.85Mn0.10Co0.05O
[111]
Mn(NO3)2·4H2O, Co(NO3)6·H2O, Zn(NO3)2·6H2O, LiOH·H2O, H2O, C2H5OH300 °C for 3 h in atmospheric airnanoparticles;
Zn0.92Co0.02Mn0.06O
[112]
Zn(Ac)2·2H2O, Mn(Ac)2·4H2O, Co(Ac)2·4H2O, NH4OH, H2O750 °C for 2 h in atmospheric airnanoparticles (15–17 nm);
Zn0.92Mn0.04Co0.04O
[113,114]
electrochemicalZnCl2, MnCl2, CoCl2, KCl, tartaric aciddensities of electrodepositioning in the range of 0.5–2.0 mA/cm2nanorod (diameters: 50–100 nm, no longer than 500 nm)[115]
Table 3. Compositions (nominal content) of precursors of Zn(1−x−y)MnxCoyO synthesis.
Table 3. Compositions (nominal content) of precursors of Zn(1−x−y)MnxCoyO synthesis.
Name of PrecursorCmZn(Ac)2∙2H2O (mol/dm3)CmMn(Ac)2∙4H2O
(mol/dm3)
CmCo(Ac)2∙4H2O
(mol/dm3)
ZnO0.303700
Zn0.98Mn0.01Co0.01O0.30370.00310.0031
Zn0.90Mn0.05Co0.05O0.30370.01690.0169
Zn0.80Mn0.10Co0.10O0.30370.03800.0380
Zn0.70Mn0.15Co0.15O0.30370.06510.0651
Table 4. Lattice parameters and ratio of lattice parameters of the obtained Zn(1−x−y)MnxCoyO NPs.
Table 4. Lattice parameters and ratio of lattice parameters of the obtained Zn(1−x−y)MnxCoyO NPs.
SampleLattice ParametersRatio of Lattice
Parameters c/a
In hcp Structure, ZnO,
Ratio of Lattice Parameters c/a
a ± σ, [Å]c ± σ, [Å]
ZnO (JCPDS No. 36-1451)3.24985.20661.60211.6330
ZnO3.25085.20741.6017
Zn0.98Mn0.01Co0.01O3.25205.20831.6010
Zn0.90Mn0.05Co0.05O3.25475.21071.6007
Zn0.80Mn0.10Co0.10O3.25635.21221.6007
Zn0.70Mn0.15Co0.15O3.25695.21341.6007
Zn0.85Mn0.15O [149]3.25645.21381.6011
Zn0.85Co0.15O [73]3.25205.20401.6003
Table 5. Results of the analysis of the chemical composition of Zn(1−x−y)MnxCoyO samples.
Table 5. Results of the analysis of the chemical composition of Zn(1−x−y)MnxCoyO samples.
SampleActual Dopant Content, mol %
EDSICP-OES
ZincManganeseCobaltZincManganeseCobalt
Zn0.98Mn0.01Co0.01O98.860.260.8899.060.250.69
Zn0.90Mn0.05Co0.05O94.361.424.2195.291.183.53
Zn0.80Mn0.10Co0.10O88.082.579.3589.942.117.96
Zn0.70Mn0.15Co0.15O83.482.8613.6585.582.3512.07
Table 6. Efficiency of ZnO codoping with Mn2+ and Co2+ ions.
Table 6. Efficiency of ZnO codoping with Mn2+ and Co2+ ions.
SampleEfficiency of Codoping Calculated Based on the Results of ICP-OES (%)
ManganeseCobalt
Zn0.98Mn0.01Co0.01O25.0069.00
Zn0.90Mn0.05Co0.05O23.6070.60
Zn0.80Mn0.10Co0.10O21.1079.60
Zn0.70Mn0.15Co0.15O15.6780.47
Table 7. Results of the colour analysis of dry NP samples (RGB and HSL colour model).
Table 7. Results of the colour analysis of dry NP samples (RGB and HSL colour model).
SampleR
(Red)
G
(Green)
B
(Blue)
H
(Hue)
S
(Saturation)
L
(Luminance)
ZnO1023102310230000001000
Zn0.85Co0.15O246376292392209304
Zn0.85Mn0.15O614438211093488403
Zn0.98Mn0.01Co0.01O660702547211194610
Zn0.90Mn0.05Co0.05O374429275226218344
Zn0.80Mn0.10Co0.10O205258145244280196
Zn0.70Mn0.15Co0.15O219160094088399152
Table 8. Characteristics of the NP samples.
Table 8. Characteristics of the NP samples.
SampleSpecific Surface Area, as ± σ (m2/g)Skeleton Density, ρs ± σ (g/cm3)Average Particle Size from SSA BET, d ± σ (nm)Average Crystallite Size from Nanopowder XRD Processor Demo, d ± σ (nm)Average crystallite Size, Scherrer’s Formula, da, dc (nm)
ZnO39.8 ± 0.15.25 ± 0.0229 ± 122 ± 720a, 24c
Zn0.98Mn0.01Co0.01O46.4 ± 0.15.24 ± 0.0224 ± 119 ± 616a, 18c
Zn0.90Mn0.05Co0.05O45.8 ± 0.15.23 ± 0.0325 ± 122 ± 1215a, 14c
Zn0.80Mn0.10Co0.10O50.6 ± 0.15.15 ± 0.0323 ± 122 ± 1314a, 11c
Zn0.70Mn0.15Co0.15O56.4 ± 0.15.06 ± 0.0221 ± 122 ± 1315a, 11c
d—diameter.

Share and Cite

MDPI and ACS Style

Wojnarowicz, J.; Omelchenko, M.; Szczytko, J.; Chudoba, T.; Gierlotka, S.; Majhofer, A.; Twardowski, A.; Lojkowski, W. Structural and Magnetic Properties of Co‒Mn Codoped ZnO Nanoparticles Obtained by Microwave Solvothermal Synthesis. Crystals 2018, 8, 410. https://doi.org/10.3390/cryst8110410

AMA Style

Wojnarowicz J, Omelchenko M, Szczytko J, Chudoba T, Gierlotka S, Majhofer A, Twardowski A, Lojkowski W. Structural and Magnetic Properties of Co‒Mn Codoped ZnO Nanoparticles Obtained by Microwave Solvothermal Synthesis. Crystals. 2018; 8(11):410. https://doi.org/10.3390/cryst8110410

Chicago/Turabian Style

Wojnarowicz, Jacek, Myroslava Omelchenko, Jacek Szczytko, Tadeusz Chudoba, Stanisław Gierlotka, Andrzej Majhofer, Andrzej Twardowski, and Witold Lojkowski. 2018. "Structural and Magnetic Properties of Co‒Mn Codoped ZnO Nanoparticles Obtained by Microwave Solvothermal Synthesis" Crystals 8, no. 11: 410. https://doi.org/10.3390/cryst8110410

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