Microwave solvothermal synthesis and characterization of manganese-doped ZnO nanoparticles

  1. 1 ,
  2. 1 ,
  3. 1,2 ,
  4. 1 ,
  5. 1 and
  6. 1
1Institute of High Pressure Physics, Polish Academy of Sciences, Sokolowska 29/37, 01-142 Warsaw, Poland
2Faculty of Materials Science and Engineering, Warsaw University of Technology, Woloska 141, 02-507 Warsaw, Poland
  1. Corresponding author email
Guest Editor: A. S. Sidorenko
Beilstein J. Nanotechnol. 2016, 7, 721–732. https://doi.org/10.3762/bjnano.7.64
Received 11 Mar 2016, Accepted 03 May 2016, Published 19 May 2016
Full Research Paper
cc by logo
Album

Abstract

Mn-doped zinc oxide nanoparticles were prepared by using the microwave solvothermal synthesis (MSS) technique. The nanoparticles were produced from a solution of zinc acetate dihydrate and manganese(II) acetate tetrahydrate using ethylene glycol as solvent. The content of Mn2+ in Zn1−xMnxO ranged from 1 to 25 mol %. The following properties of the nanostructures were investigated: skeleton density, specific surface area (SSA), phase purity (XRD), lattice parameters, dopant content, average particle size, crystallite size distribution, morphology. The average particle size of Zn1−xMnxO was determined using Scherrer’s formula, the Nanopowder XRD Processor Demo web application and by converting the specific surface area results. X-ray diffraction of synthesized samples shows a single-phase wurtzite crystal structure of ZnO without any indication of additional phases. Spherical Zn1−xMnxO particles were obtained with monocrystalline structure and average particle sizes from 17 to 30 nm depending on the content of dopant. SEM images showed an impact of the dopant concentration on the morphology of the nanoparticles.

Introduction

Nanotechnology has triggered a new global industrial revolution of the 21st century [1]. At present it is the leading technology in various research fields such as chemistry, physics, biology, medicine, materials and biomedical engineering, optoelectronics and interdisciplinary fields. It is a technology that enables testing, controlling, producing and using structures at least one dimension of which is below 100 nanometres [2]. It enables the use of nanomaterials for creating innovative products, devices and complex systems that employ the properties of materials on the nanoscale [3]. Current research of the applications of nanotechnology in optoelectronics focuses on the control of the physical properties of metal oxides semiconductors.

Zinc oxide (ZnO) is a II–VI semiconductor characterised by a wide band gap of 3.3 eV and a high exciton binding energy of circa 60 meV [4]. ZnO is used in optoelectronic devices, solar cells, data carriers, light emitting diodes (LEDs), gas sensors, thermoelectric devices, varistors, TFT display windows and laser technology [5-7]. ZnO displays pyroelectric and piezoelectric properties, thanks to which it is used in electroacoustic devices [8]. It is a biocompatible material used for producing biosensors and in drug delivery applications [9]. Thanks to antibacterial activity, matting and hiding properties, as well as bleaching properties, it is applied in the pharmaceutical and cosmetic industry to produce creams, dressings, powders, baby powders and toothpastes. In paediatric dentistry, it is the primary ingredient of the temporary filling material [10]. It is also a popular mineral filter against UVA and UVB radiation [11-13].

The search for “doped ZnO” and “Mn doped ZnO” phrases in the “ScienceDirect” scientific search engine yielded 32,757 and 7,213 matches, respectively, (April 2016) in various research fields. The number of published scientific papers confirms the growing interest in Mn2+-doped zinc oxide (Figure 1). It is generally known that selective doping enables controlling the semiconductor properties of ZnO, such as forbidden band or conductivity. The diversity of ZnO modifications by doping with various transition metal ions (e.g., Co, Mn, Ni, Fe, Cr, V) considerably increases the capabilities of applying that material in electronics, spintronics and optoelectronics [14-16].

[2190-4286-7-64-1]

Figure 1: The number of scientific publications referring to the search of “doped ZnO” and “Mn doped ZnO” phrases published in the period 2007–2015. Source: ScienceDirect (accessed April 27, 2016).

Papers related to the magnetic properties of Mn2+-doped ZnO focus on the impact of homogeneity, phase purity and dopant content on the properties [17,18]. In the case of substituting Zn2+ atoms by Mn2+ atoms in a ZnO crystalline lattice, an increase in the value of lattice parameters was observed [19,20]. The ionic radius of Mn2+ is 0.82 Å [21], which is greater than that of Zn2+ being 0.74 Å [18]. Therefore, Zn2+ can be substituted to a limited extent by Mn2+ [22]. Numerous scholars have described obtaining Mn2+-doped zinc oxide, which displays ferromagnetic properties at room temperature, in their papers [23-26]. However, there are many contradicting or even controversial publications on this topic [27-30]. Depending on the method of obtaining, Zn1−xMnxO may display either ferromagnetic [31] or paramagnetic properties [32]. Various pieces of literature reporting magnetic properties of Zn1−xMnxO NPs result from the impact of the synthesis method on: the heterogeneous distribution of Mn2+ ions in the ZnO crystalline lattice (e.g., formation of clusters [31]), the precipitation of foreign phases [33,34], the stoichiometry; the presence of oxygen vacancies [35,36], and defects of the crystalline lattice. When the synthesis or calcination of Zn1−xMnxO was carried out in an oxidising environment and the absence of foreign phases was observed, the obtained material exhibited paramagnetic properties [29,33,37]. Whereas, when Zn1−xMnxO was synthesised or calcinated for too long or at an overly high temperature, in an oxidising atmosphere, foreign phases in the form of MnO, MnO2, Mn2O, Mn3O4, as well as products of the reaction of ZnO with MnxOy such as ZnMnO3 and ZnMn2O4 were precipitated in the material. The obtained Zn1−xMnxO materials were characterised by either ferromagnetic or paramagnetic properties depending on the presence of foreign phases [33,36,38,39]. This results from different magnetic properties of MnxOy and metallic Mn. Metallic manganese is antiferromagnetic, while many alloys of manganese, in which the average Mn–Mn distance is greater than that of metallic manganese, are ferromagnetic [40]. MnO, Mn2O3 and MnO2 are antiferromagnetic [41], while Mn3O4 is ferromagnetic [42,43]. ZnMnO3 is paramagnetic [44,45]. But when a reducing environment was selected for sample calcination, also precipitations of ZnMnO3 and ZnMn2O4 phase emerged in Zn1−xMnxO and the material displayed only ferromagnetic properties [36]. Ferromagnetic properties of Zn1−xMnxO are explained by the presence of oxygen vacancies, manganese in various oxidation states (II–IV), and dopant clusters. The formation of clusters of Mn2+ dopant in ZnO NPs, e.g., on the particle surface, may lead to a various kind of magnetic coupling of atoms with their adjacent atoms [46]. This, in turn, leads to different magnetic properties than those encountered in nanomaterials with the same dopant content but with a different distribution of ions or dopant clusters in the crystalline lattice (Figure 2) [47].

[2190-4286-7-64-2]

Figure 2: Indicative examples of the possible distribution of Mn2+ dopant in zinc oxide crystalline lattice. Black circles: Mn2+ dopant, white circles: ZnO. a) Unclustered distribution of dopant; b) clustered distribution of dopant; c) random distribution of dopant.

Despite the efforts made by numerous research groups and despite the development of new synthesis methods, obtaining nanomaterials with reproducible properties proves a very complex and problematic task. Obtaining Zn1−xMnxO nanomaterials with reproducible optoelectronic and magnetic properties remains an unsolved issue. Differing properties of the obtained Zn1−xMnxO result from the complexity of chemical reactions and the limitations of the currently employed methods. The lack of simultaneous control over chemical composition, stoichiometry, dopant homogeneity, particle size distribution, shape, phase purity, surface modification and agglomeration, makes it difficult to obtain NPs [22]. The primary cause of the lack of reproducibility of magnetic properties in the case of Zn1−xMnxO is the lack of control over the doping impact on the formation of oxygen vacancies, crystalline lattice defects and dopant clusters [48]. It is presumed that there are several competitive chemical reactions in the reaction of Zn1−xMnxO synthesis, such as the oxidisation of Mn2+ ions, the formation of MnxOy, the reduction of Mn2+ ions to metallic Mn, and the reaction of MnxOy with ZnO leading to the formation of spinels (ZnMnO3, ZnMn2O4). These reactions may occur simultaneously depending on the synthesis method, the synthesis parameters and the method of precursor preparation. This explains the various reports about obtaining Zn1−xMnxO with different and in parts contradictory properties. The relevant literature describes several methods for the synthesis of Zn1−xMnxO nanomaterials, in particular coprecipitation [49], coprecipitation–calcination [29,50,51], combustion [52], sol–gel [7,31,33,53,54], spin coating [19,20], pulsed laser deposition [26], reactions in the solid state [35,55], thermal evaporation [56], and thermal evaporation vapour-phase deposition [30]. However, two of the most promising methods, which are still being developed are hydrothermal and solvothermal synthesis [16,57]. The growing popularity of these methods is proved by the emergence of new types of reactors, e.g., microwave stop-flow and continuous-flow in supercritical water [57-63]. Microwave solvothermal synthesis (MSS) is quicker, purer, and more energy and cost efficient than conventional synthesis methods. The microwave radiation employed is a highly effective method of providing energy to the reaction chamber, which results in a more uniform and rapid heating in comparison with traditional methods of heat transfer [64]. The MSS method is a “wet chemistry” method. This means that the precursor is obtained from solutions of substrates, which are mixed in precisely determined quantities, and the synthesis is carried out under controlled conditions (temperature, pressure, time). The synthesis product is a precipitate, which is mainly subjected to filtering, rinsing and drying [65]. The advantages of the microwave solvothermal synthesis are purity (Teflon reaction chamber, contactless heating method), short process duration, precise control over parameters (time, pressure), product homogeneity (volumetric heating), control over dopant content, high efficiency, and surface modification [47]. MSS products are characterised by homogeneous morphology, purity, narrow size distribution and low agglomeration.

The present paper contains an attempt to obtain ZnO nanoparticles with a Mn2+ dopant content of up to 25 mol %. For the synthesis we selected an organic solvent with weak reducing properties to prevent the precipitation of foreign phases, which is caused above all by the changed oxidisation state of Mn2+ ions. The microwave solvothermal synthesis was selected for obtaining Zn1−xMnxO. It was shown before by us that the MSS method led to fully crystalized Co2+-doped ZnO without foreign phases and with a uniform particle size and shape [66].

Experimental

Substrates

Zinc acetate dihydrate (Zn(CH3COO)2·2H2O), analytically pure, SKU: 112654906-1KG; ethylene glycol (1,2-ethanediol, C2H4(OH)2), pure, SKU: 114466303-5L and manganese(II) acetate tetrahydrate (Mn(CH3COO)2·4H2O), analytically pure, SKU: 116167509-100g were purchased from Chempur. The reagents were used without additional purification.

Synthesis of Zn1−xMnxO NPs

Zn1−xMnxO NPs were obtained by the MSS method. The precursor was obtained by dissolving powder mixtures, composed of 1, 5, 10, 15, 20, 25 mol % of manganese acetate in zinc acetate, in ethylene glycol (Table 1). The reference sample of ZnO was obtained from the precursor without adding the Mn2+ doping. The mixture of the acetates in glycol (75 mL) was heated to 70 °C and stirred using a magnetic stirrer (SLR, SI Analytics, Germany) until the components were completely dissolved. After cooling down to ambient temperature the solution was poured to the Teflon reaction chamber with a total volume of 110 mL [66].

Table 1: Composition of precursors of Zn1−xMnxO synthesis.

name of precursor c(Zn(CH3COO)2·2H2O) [mol/L] c(Mn(CH3COO)2∙4H2O) [mol/L] solvent
ZnO 0.3254 0 ethylene glycol
Zn0.99Mn0.01O 0.3254 0.0033
Zn0.95Mn0.05O 0.3254 0.0171
Zn0.90Mn0.10O 0.3254 0.0362
Zn0.85Mn0.15O 0.3254 0.0574
Zn0.80Mn0.20O 0.3254 0.0814
Zn0.75Mn0.25O 0.3254 0.1085

The synthesis reaction initiated by microwave radiation was carried out in a microwave reactor (Model 02-02, 600 W, 2.45 GHz, ERTEC, Poland) at a temperature of 200 °C. The synthesis of Mn2+-doped zinc oxide in ethylene glycol can be described through the following equation:

[Graphic 1]

The reaction duration was 25 min and afterwards the reactor chamber was cooled down for 20 min to reach a temperature below 100 °C in order to safely remove the reaction vessel. After the synthesis the suspension was subjected to sedimentation and decantation. The sediment was rinsed three times with deionised water (HLP 20UV, Hydrolab, Poland). In order to obtain a dry Zn1−xMnxO powder, the sediment was suspended in water and rapidly cooled down with liquid nitrogen and subsequently dried in a freeze dryer (Lyovac GT-2, SRK Systemtechnik GmbH, Germany).

Characterisation methods

X-ray powder diffraction

Diffraction patterns of the X-ray powder diffraction (XRD) were gathered at room temperature in the 2 theta range from 10 to 100° with a step width of 0.02° (Cu Kα1, X’Pert PRO, Panalytical, Netherlands). The crystalline lattice parameters were determined by the Rietveld method. Diffraction patterns were fitted in Fityk software, version 0.9.8, using the implemented Pearson 7 feature. Based on the diffraction patterns, the size of crystallites was determined (Dhkl) using Scherrer’s formula (Equation 1), where Dhkl is the volume weighted crystallite size [nm]; β is the FWHM of the hkl diffraction peak [rad]; K is a constant shape factor (K ≈ 1), λ is the X-ray wave length [nm] and θhkl is the Bragg diffraction angle.

[2190-4286-7-64-i1]
(1)

Crystallite size distribution

The analysis of XRD peak profiles was performed using the analytical formula for polydispersive powders [67]. While the Scherrer method provides a single size parameter, this technique provides four parameters: average crystallite size, error of the average crystallite size, dispersion of size and error of dispersion of sizes. Hence, a full crystallite size distribution curve with its errors is obtained. The on-line tool Nanopowder XRD Processor Demo (http://science24.com/xrd/) is a webpage where diffraction files can be uploaded. Files are processed on a server to extract the crystallite size distribution for XRD peaks. Unlike the standard fitting, the tool does not act in the reciprocal space at all, but solves sets of equations in a few auxiliary spaces simultaneously. This allows for an analysis of XRD data with heavily convoluted reciprocal space peaks [68,69].

Measurement of density and specific surface area

Density measurements were carried out using a helium pycnometer (AccuPyc II 1340 FoamPyc V1.06, Micromeritics, USA), in accordance with ISO 12154:2014 at the temperature of 25 ± 2 °C. The density of a material obtained using helium pycnometry is called, e.g., skeleton density, pycnometric density, true density and helium density in the relevant literature. The specific surface area of NPs was determined using the analysis of nitrogen adsorption isotherm by the BET (Brunauer–Emmett–Teller) method (Gemini 2360, V 2.01, Micromeritics), in accordance with ISO 9277:2010. Prior to performing measurements of both density and specific surface area, the samples were subject to 2 h desorption in a desorption station (FlowPrep, 060 Micromeritics), at a temperature of 150 °C with a flow of helium of 99.999% purity. Based on the determined specific surface area and density, the average size of particles was determined, with the assumption that all particles are spherical and identical [70]. Equation 2 was used for calculating the average particle size, where D is the average diameter of the of particles [µm], N is a shape coefficient being (N = 6) for sphere, SSA is the specific surface area [m2/g] and ρ the density [g/cm3].

[2190-4286-7-64-i2]
(2)

Morphologic characteristics

The morphology of NPs was determined using scanning electron microscopy (SEM) (ZEISS, Ultra Plus, Germany). The samples were coated with a thin carbon layer using a sputter coater (SCD 005/CEA 035, BAL-TEC, Switzerland).

Chemical composition analysis

The chemical composition analysis was carried out by using visible-light spectrophotometry (DR 3900, Hach Lange, Germany). The analysis of the content of Zn2+ ions was carried out by using the Zincon method No. 8009 (Hach Lange). The analysis of the content of Mn2+ ions was carried out by using the 1-(2-pyridylazo)-2-naphthol (PAN) method No. 8149 (Hach Lange). The chemical composition analysis was carried out by using inductively coupled argon plasma optical emission spectrometry (ICP-OES) (Thermo Scientific, iCAP model 6000, Great Britain).

Samples for the quantitative analysis were prepared as follows: 5 mg of powder was weighed in a 110 mL Teflon vessel and 15 mL of deionised water (HLP 20UV, Hydrolab, Poland) was added; then 6 mL of HNO3 was added and the solution subjected to one microwave heating cycle in a Magnum II reactor (600 W, 2.45 GHz, ERTEC, Poland). After cooling the sample volume was filled up to 50 mL with deionized water [66]. The pH of samples before the colorimetric analysis was adjusted in accordance with the recommendations of procedures No. 8009 and No. 8149.

Quantitative microanalysis was carried out through energy-dispersive X-ray spectroscopy (EDS) by using an EDS analyser (Quantax 400, Bruker, USA). Samples for EDS tests were pressed to pellets with a diameter of 5 mm.

Results and Discussion

Morphology

Figure 3 and Figure 4 present representative SEM images of Zn1−xMnxO NPs. They show particles with homogeneously distributed spherical shapes. An impact of the content of Mn2+ dopant on the morphology of Zn1−xMnxO powder was observed. Powders with dopant contents of 5, 10, 15, 20 and 25 mol % are composed of dense structures forming conglomerates whose shape resembles the structure of a cauliflower (Figure 4). In the case of dopant contents of 0 or 1 mol %, loose nanopowders were obtained. It can be noticed that the average particle size decreases with increasing content of Mn2+ dopant in the NPs. The average particle size decreases from circa 30–35 nm to 15–25 nm with an increase in the content of Mn2+ dopant from 0 to 20 mol %. In the case of the powder with a dopant content of 25 mol %, the average particle size was circa 20–25nm. SEM tests reveal that each precursor composition must be treated individually. That means that synthesis parameters must be optimised for each precursor composition in order to eliminate the process of agglomeration and formation of conglomerates of Zn1−xMnxO NPs.

[2190-4286-7-64-3]

Figure 3: SEM images of NPs of Zn1−xMnxO with dopant concentrations of: a) 0 mol %, b) 1 mol %, c) 5 mol %, d) 10 mol %, e) 15 mol %, f) 20 mol % and g) 25 mol %.

[2190-4286-7-64-4]

Figure 4: SEM images of agglomerates and conglomerates of Zn1−xMnxO NPs with dopant concentrations of: a) 0 mol %, b) 1 mol %, c) 5 mol %, d) 10 mol %, e) 15 mol %, f) 20 mol % and g) 25 mol %.

Phase composition

The XRD tests of the samples did not reveal a presence of foreign phases in the obtained Zn1−xMnxO NPs, and all diffraction peaks can be attributed to the hexagonal phase ZnO (Figure 5). Zinc oxide is characterised by hexagonal wurtzite structure (JCPDS No. 36-1451, space group: P63mc) with two lattice parameters a and c [4]. The parameters of ZnO crystalline lattice assume the values a = 3.2498 Å and c = 5.2066 Å, where its c/a ratio of 1.6021 is close to that of a close-packed hexagonal structure (hcp) c/a = 1.6330. MnO has rock salt structure (space group: Fm−3m) with the lattice parameter a = 4.4475 Å [71]. The radius of Mn2+ ions is bigger than that of Zn2+ ions by 0.08 Å, which explains the considerable changes in the lattice parameters (Table 2). Both lattice parameters a and c increase with increasing dopant content from 1 to 25 mol % (Figure 6). The results of the c/a lattice parameter ratio reveals that the change of the dopant content in Zn1−xMnxO leads to a change in the proportions of the unit cell dimensions (Table 2). The obtained results of crystallite sizes da and dc (Table 3, Figure 7) indicate that the increase of Mn2+-dopant content in ZnO leads to changes in proportions (asymmetry) of crystallite sizes. In line with the increase in Mn2+-dopant content, the size of the crystallites in the c-direction decreases (Figure 7). The dc/da ratio can be interpreted as a change in the shape of Zn1−xMnxO particles only if they are monocrystalline, which is the case for our samples (Table 3). If the dc/da ratio assumes the value of 1, the shape of the Zn1−xMnxO particles is spherical. If the dc/da ratio assumes a value of different from 1, means particles exhibit a spherical/elliptical shape. Despite the presence of as much as 25 mol % dopant in the precursor solution, no foreign phases in the form of ZnMnO3, ZnMn2O4, metallic Mn, or MnxOy were observed in the synthesis product, which proves that the organic solvent selected by us (ethylene glycol) prevented competitive reactions and a change of the oxidisation state of Mn2+ in the course of the synthesis. Nevertheless, it should be remembered that the limit of foreign phase detectability in the diffraction method (XRD) can be as high as 5–6 atom %.

[2190-4286-7-64-5]

Figure 5: XRD diffraction patterns of Zn1−xMnxO NPs, with the nominal dopant content in the solution being 0, 1, 5, 10, 15, 20, 25 mol % and its comparison with the standard pattern of wurtzite-phase ZnO (JCPDS No. 36-1451).

Table 2: Lattice parameters and ratio of lattice parameters of the obtained Zn1−xMnxO NPs.

sample lattice parameters lattice parameter ratio c/a lattice parameter ratio c/a in hcp ZnO
a ± σ [Å] c ± σ [Å]
ZnO (JCPDS No. 36-1451) 3.2498 5.2066 1.6021 1.6330
ZnO 3.2499 ± 0.0001 5.2066 ± 0.0003 1.6021
Zn0.99Mn0.01O 3.2506 ± 0.0002 5.2072 ± 0.0004 1.6019
Zn0.95Mn0.05O 3.2523 ± 0.0003 5.2092 ± 0.0006 1.6017
Zn0.90Mn0.10O 3.2535 ± 0.0004 5.2111 ± 0.0004 1.6017
Zn0.85Mn0.15O 3.2564 ± 0.0005 5.2138 ± 0.0008 1.6011
Zn0.80Mn0.20O 3.2564 ± 0.0004 5.2146 ± 0.0004 1.6013
Zn0.75Mn0.25O 3.2593 ± 0.0005 5.2174 ± 0.0008 1.6008
[2190-4286-7-64-6]

Figure 6: Lattice parameters versus nominal Mn2+ content of Zn1−xMnxO samples.

Table 3: Crystallite sizes and size ratios of the obtained Zn1−xMnxO NPs.

sample average crystallite size, calculated through Scherrer’s formula size ratio dc/da
da [nm] dc [nm]
ZnO 22 28 1.273
Zn0.99Mn0.01O 21 26 1.238
Zn0.95Mn0.05O 18 19 1.056
Zn0.90Mn0.10O 17 15 0.882
Zn0.85Mn0.15O 18 15 0.833
Zn0.80Mn0.20O 16 12 0.750
Zn0.75Mn0.25O 19 13 0.684
[2190-4286-7-64-7]

Figure 7: Dependence of the changes in crystallite size on the nominal content of Mn2+-dopant in Zn1−xMnxO.

Chemical composition and impact of dopant on the colour of NPs

The real content (RC) of manganese dopant in Zn1−xMnxO is presented in Table 4. The RC of dopant in Zn1−xMnxO is smaller by circa 78% than the expected value of dopant arising from the nominal content (NC) of the dopant in the prepared precursor solution. The difference in sizes of Zn2+ and Mn2+ ionic radii is the reason for the low doping efficiency of Zn1−xMnxO NPs. The chemical composition analysis was carried out by three methods: ICP-OES, colorimetry and X-ray microanalysis. The most often used methods in the relevant literature are ICP-OES and X-ray microanalysis. All the employed methods provided mutually different results (Table 4). This resulted from the limitations and accuracy of the various analytical techniques in relation to the quantitative determination of zinc and manganese. ICP-OES analysis is regarded as the most accurate and precise method. Figure 8 describes the correlation between NC in the precursor and RC of Mn2+ dopant in NPs. The average efficiency of doping during the synthesis of Zn1−xMnxO in ethylene glycol was merely 22 mol %, being four times smaller than in the case of doping ZnO with Co2+ ions [66]. The derived formula (y = −0.0057x2 + 0.353x − 0.0447, where y is the real and x is the nominal dopant content) enables the control of RC of the dopant in Zn1−xMnxO NPs at the stage of precursor preparation. The nominal contents of manganese dopant in Zn1−xMnxO were used for denominating the samples in this paper.

Table 4: Analysis of chemical composition of Zn1−xMnxO samples.

name of sample actual content of dopant, mol %
colorimetric analysis [mol %] EDS ICP-OES
zinc manganese zinc manganese zinc manganese
Zn0.99Mn0.01O 99.41 0.59 99.86 0.14 99.70 0.30
Zn0.95Mn0.05O 98.63 1.37 98.73 1.27 98.47 1.53
Zn0.90Mn0.10O 97.74 2.26 97.58 2.42 97.23 2.77
Zn0.85Mn0.15O 97.04 2.96 96.06 3.78 95.45 4.29
Zn0.80Mn0.20O 96.53 3.47 96.22 3.94 95.71 4.55
Zn0.75Mn0.25O 96.03 3.97 93.89 6.11 94.72 5.28
[2190-4286-7-64-8]

Figure 8: Correlation between the nominal content and the real content of Mn2+ dopant in ZnO. The results were obtained based on the ICP-OES analysis.

The doping of zinc oxide with Mn2+ ions leads to a change in the optical properties of the material, which is proved by the change of the sample colour (Figure 9 and Figure 10). The obtained Zn1−xMnxO NPs, depending on the Mn2+-dopant content, have differing shades of yellow and orange. The intensity of NPs colour depends on the dopant content, which is illustrated in Figure 9 and Figure 10. Most of the Zn1−xMnxO nanostructure materials exhibit strong absorption in the UV and visible range and a strong UV emission [30].

[2190-4286-7-64-9]

Figure 9: Visual comparison of changes in colours of suspensions of Zn1−xMnxO NPs depending on the dopant content.

[2190-4286-7-64-10]

Figure 10: Visual comparison of changes in colours of dry powder of Zn1−xMnxO NPs depending on the dopant content.

Density, specific surface area and crystallite size distribution

The theoretical density of ZnO is 5.606 g/cm3. The pycnometric density of the obtained reference sample of ZnO NPs without doping amounted to 5.09 g/cm3. The difference between the nano density and the theoretical density of ZnO can result above all from surface defects, the presence of Zn(OH)2 hydroxides and the non-stoichiometric composition. The density of Zn1−xMnxO, irrespectively of the content of Mn2+ dopant, was constant at about 5.0 g/cm3. The lower density of Zn1−xMnxO is related to the smaller atomic weight of Mn (54.93 u) in comparison with the Zn atoms (65.38 u) in Zn1−xMnxO, and to the possible presence of a greater quantity of defects in the crystalline lattice. The specific surface area of Zn1−xMnxO ranged from 40 to 70 m2/g (Table 5). The average size of Zn1−xMnxO particles calculated based on the specific surface area ranged from 17 to 30nm (Table 5). Based on the XRD analysis using the Nanopowder XRD Processor Demo web application, the average size and the size distribution of crystallites were determined for Zn1−xMnxO NPs (Figure 11). Zn1−xMnxO material was characterised by an average crystallite size from 19 to 26 nm (Table 5), with a narrow distribution, which ranged from 30–35 nm to 40–60 nm depending on the dopant content (Figure 11). When comparing the obtained results from the Nanopowder XRD Processor Demo application and with that from Scherrer’s formula, similar crystallite sizes were obtained. The results fall within the range of the standard deviation of methods (Table 5). The calculations of the average crystallite size (XRD) coincide with the average particle sizes calculated based on the specific surface are with an accuracy of 2–6 nm, which confirms that:

  • the nanoparticles are monocrystalline;
  • the formed structures of nanoparticle conglomerates (Figure 2 and Figure 3) do not form aggregates;
  • the size of crystallites is equal to the size of nanoparticles, thanks to which these expressions can be used interchangeably in this paper;
  • the obtained crystallite size distribution may be interpreted as particle size distribution.

Table 5: Characteristic of the Zn1−xMnxO NPs samples.

sample specific surface area by gas adsorption, as ± σ [m2/g] skeleton density by gas pycnometry, 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 from Scherrer’s formula
          da [nm] dc [nm]
ZnO 42±1 5.09±0.04 28±1 26±8 22 28
Zn0.99Mn0.01O 40±1 4.98±0.05 30±1 24±8 21 26
Zn0.95Mn0.05O 50±1 4.95±0.03 24±1 21±6 18 19
Zn0.90Mn0.10O 60±1 5.03±0.04 20±1 19±6 17 15
Zn0.85Mn0.15O 57±1 5.01±0.04 21±1 20±7 18 15
Zn0.80Mn0.20O 70±1 4.97±0.03 17±1 19±10 16 12
Zn0.75Mn0.25O 63±1 4.97±0.04 19±1 21±9 19 13
[2190-4286-7-64-11]

Figure 11: Crystallite size distribution obtained using Nanopowder XRD Processor Demo [68].

Conclusion

Zn1−xMnxO (x = 0.01, 0.05, 0.10, 0.15, 0.2, 0.25) nanoparticles have been synthesized by microwave solvothermal synthesis. Zinc and manganese acetates dissolved in ethylene glycol were used as precursor. Crystalline Zn1−xMnxO NPs with hexagonal wurtzite structure, pure in terms of phase, were obtained. Zn1−xMnxO was characterised by an average particle size between 17 and 30 nm, a specific surface area between 40 and 70 m2/g and a density of circa 5.0 g/cm3. The average crystallite size was observed to decrease in line with an increase in the Mn2+ content. SEM images did not reveal an impact of the dopant on the spherical shape of nanoparticles. However, an impact of an increase in the Mn2+ dopant content on the shape of agglomerates and conglomerates of Zn1−xMnxO nanoparticles was observed. The crystalline lattice parameters a and c in Zn1−xMnxO increase in line with the dopant growth, which is a consequence of the greater ionic radius of Mn2+ dopant. The results confirm the usability of the developed method for obtaining crystalline Zn1−xMnxO, pure in terms of phase, with a narrow particle size distribution with the nominal dopant content reaching 25 mol % and doping efficiency of circa 22%. Our paper also indicates a high potential of microwave solvothermal synthesis for obtaining doped ZnO nanoparticles.

Acknowledgements

The work reported here was conducted in the Shyman Project, grant agreement no. 280983 (2012–2016) under the 7th Framework Programme of the European Commission, coordinated by E. Lester (The University of Nottingham, United Kingdom, http://www.shyman.eu). The research subject was carried out with the use of equipment funded by the project CePT, reference: POIG.02.02.00-14-024/08, financed by the European Regional Development Fund within the Operational Programme “Innovative Economy” for 2007–2013 (http://www.cept.wum.edu.pl).

References

  1. Foresight review of nanotechnology, The next industrial revolution; Lloyd’s Register Foundation Report Series: No.2014.1, April 2014.
    http://www.lrfoundation.org.uk/news/2014/nanotechnology-review.aspx
    Return to citation in text: [1]
  2. Nanotechnologies -- Vocabulary -- Part 2: Nano-objects; ISO/TS 80004-2:2015.
    Return to citation in text: [1]
  3. Vance, M. E.; Kuiken, T.; Vejerano, E. P.; McGinnis, S. P.; Hochella, M. F., Jr.; Rejeski, D.; Hull, M. S. Beilstein J. Nanotechnol. 2015, 6, 1769–1780. doi:10.3762/bjnano.6.181
    Return to citation in text: [1]
  4. Morkoç, H.; Özgür, Ü. Zinc Oxide: Fundamentals, Materials and Device Technology; Wiley-VCH: Heidelberg, Germany, 2009. doi:10.1002/9783527623945
    Return to citation in text: [1] [2]
  5. Özgür, Ü.; Alivov, Ya. I.; Liu, C.; Teke, A.; Reshchikov, M. A.; Doğan, S.; Avrutin, V.; Cho, S.-J.; Morkoç, H. J. Appl. Phys. 2005, 98, 041301. doi:10.1063/1.1992666
    Return to citation in text: [1]
  6. Ozgur, U.; Hofstetter, D.; Morkoc, H. Proc. IEEE 2010, 98, 1255–1268. doi:10.1109/JPROC.2010.2044550
    Return to citation in text: [1]
  7. Dilonardo, E.; Penza, M.; Alvisi, M.; Di Franco, C.; Palmisano, F.; Torsi, L.; Cioffi, N. Beilstein J. Nanotechnol. 2016, 7, 22–31. doi:10.3762/bjnano.7.3
    Return to citation in text: [1] [2]
  8. Chen, J.; He, X.; Wang, W.; Xuan, W.; Zhou, J.; Wang, X.; Dong, S. R.; Garner, S.; Cimo, P.; Luo, J. K. J. Mater. Chem. C 2014, 2, 9109–9114. doi:10.1039/C4TC01307G
    Return to citation in text: [1]
  9. Xiong, H.-M. Adv. Mater. 2013, 25, 5329–5335. doi:10.1002/adma.201301732
    Return to citation in text: [1]
  10. Strange, D. M.; Strange, M.; Sue Scale, N.; Nunn, M. E. Pediatr. Dent. 2001, 23, 331–336.
    Return to citation in text: [1]
  11. Pinnell, S. R.; Fairhurst, D.; Gillies, R.; Mitchnick, M. A.; Kollias, N. Dermatol. Surg. 2000, 26, 309–314. doi:10.1046/j.1524-4725.2000.99237.x
    Return to citation in text: [1]
  12. Sun, Q.-C.; Xu, X.; Musfeldt, J. L.; Baker, S. N.; Christianson, A. D. Chem. Mater. 2011, 23, 2956–2960. doi:10.1021/cm200582t
    Return to citation in text: [1]
  13. Mitchnick, M. A.; Fairhurst, D.; Pinnell, S. R. J. Am. Acad. Dermatol. 1999, 40, 85–90. doi:10.1016/S0190-9622(99)70532-3
    Return to citation in text: [1]
  14. Ohno, H. Science 1998, 281, 951–956. doi:10.1126/science.281.5379.951
    Return to citation in text: [1]
  15. Pearton, S. J.; Norton, D. P.; Ivill, M. P.; Hebard, A. F.; Chen, W. M.; Buyanova, I. A.; Zavada, J. M. J. Electron. Mater. 2007, 999, 0999-K03-04. doi:10.1557/PROC-0999-K03-04
    Return to citation in text: [1]
  16. Glaspell, G.; Dutta, P.; Manivannan, A. J. Cluster Sci. 2005, 16, 523–536. doi:10.1007/s10876-005-0024-y
    Return to citation in text: [1] [2]
  17. Jing, C.; Jiang, Y.; Bai, W.; Chu, J.; Liu, A. J. Magn. Magn. Mater. 2010, 322, 2395–2400. doi:10.1016/j.jmmm.2010.02.044
    Return to citation in text: [1]
  18. Hao, Y.-M.; Lou, S.-Y.; Zhou, S.-M.; Yuan, R.-J.; Zhu, G.-Y.; Li, N. Nanoscale Res. Lett. 2012, 7, No. 100. doi:10.1186/1556-276X-7-100
    Return to citation in text: [1] [2]
  19. Ilyas, U.; Rawat, R. S.; Wang, Y.; Tan, T. L.; Lee, P.; Chen, R.; Sun, H. D.; Li, F.; Zhang, S. Appl. Surf. Sci. 2012, 258, 6373–6378. doi:10.1016/j.apsusc.2012.03.043
    Return to citation in text: [1] [2]
  20. Ali Fatima, A.; Devadason, S.; Mahalingam, T. J. Mater. Sci.: Mater. Electron. 2014, 25, 3466–3472. doi:10.1007/s10854-014-2040-x
    Return to citation in text: [1] [2]
  21. Allaby, M., Ed. A Dictionary of Earth Sciences, 3rd ed.; Oxford University Press: Oxford, United Kingdom, 2008. doi:10.1093/acref/9780199211944.001.0001
    Return to citation in text: [1]
  22. Lojkowski, W.; Gedanken, A.; Grzanka, E.; Opalinska, A.; Strachowski, T.; Pielaszek, R.; Tomaszewska-Grzeda, A.; Yatsunenko, S.; Godlewski, M.; Matysiak, H.; Kurzydłowski, K. J. J. Nanopart. Res. 2009, 11, 1991–2002. doi:10.1007/s11051-008-9559-9
    Return to citation in text: [1] [2]
  23. Pradhan, A. K.; Kai, Z.; Mohanty, S.; Dadson, J. B.; Hunter, D.; Zhang, J.; Sellmyer, D. J.; Roy, U. N.; Cui, Y.; Burger, A.; Mathews, S.; Joseph, B.; Sekhar, B. R.; Roul, B. K. Appl. Phys. Lett. 2005, 86, 152511. doi:10.1063/1.1897827
    Return to citation in text: [1]
  24. Chang, Y. Q.; Wang, D. B.; Luo, X. H.; Xu, X. Y.; Chen, X. H.; Li, L.; Chen, C. P.; Wang, R. M.; Xu, J.; Yu, D. P. Appl. Phys. Lett. 2003, 83, 4020–4022. doi:10.1063/1.1625788
    Return to citation in text: [1]
  25. Kundaliya, D. C.; Ogale, S. B.; Lofland, S. E.; Dhar, S.; Metting, C. J.; Shinde, S. R.; Ma, Z.; Varughese, B.; Ramanujachary, K. V.; Salamanca-Riba, L.; Venkatesan, T. Nat. Mater. 2004, 3, 709–714. doi:10.1038/nmat1221
    Return to citation in text: [1]
  26. Aravind, A.; Jayaraj, M. K.; Kumar, M.; Chandra, R. Mater. Sci. Eng., B 2012, 177, 1017–1022. doi:10.1016/j.mseb.2012.05.005
    Return to citation in text: [1] [2]
  27. Fukumura, T.; Jin, Z.; Kawasaki, M.; Shono, T.; Hasegawa, T.; Koshihara, S.; Koinuma, H. Appl. Phys. Lett. 2001, 78, 958–960. doi:10.1063/1.1348323
    Return to citation in text: [1]
  28. Kim, S. S.; Moon, J. H.; Lee, B.-T.; Song, O. S.; Je, J. H. J. Appl. Phys. 2004, 95, 454–459. doi:10.1063/1.1632547
    Return to citation in text: [1]
  29. Alaria, J.; Bouloudenine, M.; Schmerber, G.; Colis, S.; Dinia, A.; Turek, P.; Bernard, M. J. Appl. Phys. 2006, 99, 08M118. doi:10.1063/1.2172887
    Return to citation in text: [1] [2] [3]
  30. Ma, X.; Wang, Z. Microelectron. Eng. 2001, 88, 3168–3171. doi:10.1016/j.mee.2011.06.021
    Return to citation in text: [1] [2] [3]
  31. Li, J. H.; Shen, D. Z.; Zhang, J. Y.; Zhao, D. X.; Li, B. S.; Lu, Y. M.; Liu, Y. C.; Fan, X. W. J. Magn. Magn. Mater. 2006, 302, 118–121. doi:10.1016/j.jmmm.2005.08.025
    Return to citation in text: [1] [2] [3]
  32. Sharma, V. K.; Varma, G. D. J. Alloys Compd. 2008, 458, 523–527. doi:10.1016/j.jallcom.2007.04.033
    Return to citation in text: [1]
  33. Blasco, J.; Bartolomé, F.; García, L. M.; García, J. J. Mater. Chem. 2006, 16, 2282–2288. doi:10.1039/b518418e
    Return to citation in text: [1] [2] [3] [4]
  34. Zhang, B.; Li, M.; Wang, J. Z.; Shi, L. Q.; Cheng, H. S. Mater. Sci. Appl. 2013, 4, 307–311. doi:10.4236/msa.2013.45039
    Return to citation in text: [1]
  35. Chattopadhyay, S.; Neogi, S. K.; Sarkar, A.; Mukadam, M. D.; Yusuf, S. M.; Banerjee, A.; Bandyopadhyay, S. J. Magn. Magn. Mater. 2011, 323, 363–368. doi:10.1016/j.jmmm.2010.09.042
    Return to citation in text: [1] [2]
  36. Sharma, V. K.; Varma, G. D. J. Appl. Phys. 2007, 102, 056105. doi:10.1063/1.2778283
    Return to citation in text: [1] [2] [3]
  37. Tuyen, N. V.; Canh, T. D.; Long, N. N.; Nghia, N. X.; Trinh, B. N. Q.; Shen, Z. J. Phys.: Conf. Ser. 2009, 187, 012020. doi:10.1088/1742-6596/187/1/012020
    Return to citation in text: [1]
  38. Toloman, D.; Mesaros, A.; Popa, A.; Raita, O.; Silipas, T. D.; Vasile, B. S.; Pana, O.; Giurgiu, L. M. J. Alloys Compd. 2013, 551, 502–507. doi:10.1016/j.jallcom.2012.10.183
    Return to citation in text: [1]
  39. Song, Y. Y.; Park, K. S.; Park, K. H.; Oh, S. K.; Yu, S. C.; Kang, H. J.; Lee, K. W. J. Korean Phys. Soc. 2008, 52, 106–111. doi:10.3938/jkps.52.106
    Return to citation in text: [1]
  40. Eringen, A. C.; Maugin, G. A. Electrodynamics of Continua I: Foundations and Solid Media; Springer: Berlin, Germany, 1990. doi:10.1007/978-1-4612-3226-1
    Return to citation in text: [1]
  41. Jung, S. W.; An, S.-J.; Yi, G.-C.; Jung, C. U.; Lee, S.-I.; Cho, S. Appl. Phys. Lett. 2002, 80, 4561. doi:10.1063/1.1487927
    Return to citation in text: [1]
  42. Liu, C.; Yun, F.; Morkoç, H. J. Mater. Sci.: Mater. Electron. 2005, 16, 555–597. doi:10.1007/s10854-005-3232-1
    Return to citation in text: [1]
  43. Yang, L.-X.; Zhu, Y.-J.; Tong, H.; Wang, W.-W.; Cheng, G.-F. J. Solid State Chem. 2006, 179, 1225–1229. doi:10.1016/j.jssc.2006.01.033
    Return to citation in text: [1]
  44. Jaćimović, J.; Micković, Z.; Gaál, R.; Smajda, R.; Vâju, C.; Sienkiewicz, A.; Forró, L.; Magrez, A. Solid State Commun. 2011, 151, 487–490. doi:10.1016/j.ssc.2010.12.025
    Return to citation in text: [1]
  45. Blasco, J.; García, J. J. Solid State Chem. 2006, 179, 2199–2205. doi:10.1016/j.jssc.2006.04.026
    Return to citation in text: [1]
  46. Saal, H.; Binnewies, M.; Bredow, T. Phys. Chem. Chem. Phys. 2009, 11, 3201–3209. doi:10.1039/b901596e
    Return to citation in text: [1]
  47. Wojnarowicz, J.; Kusnieruk, S.; Chudoba, T.; Mizeracki, J.; Łojkowski, W. Glass Ceram. 2015, 3, 8–13.
    Return to citation in text: [1] [2]
  48. Senthilkumaar, S.; Rajendran, K.; Banerjee, S.; Chini, T. K.; Sengodan, V. Mater. Sci. Semicond. Process. 2008, 11, 6–12. doi:10.1016/j.mssp.2008.04.005
    Return to citation in text: [1]
  49. Sharma, R. K.; Patel, S.; Pargaien, K. C. Adv. Nat. Sci.: Nanosci. Nanotechnol. 2012, 3, 035005. doi:10.1088/2043-6262/3/3/035005
    Return to citation in text: [1]
  50. Abdollahi, Y.; Abdullah, A. H.; Zainal, Z.; Yusof, N. A. Int. J. Basic Appl. Sci. 2011, 11, 44–50.
    Return to citation in text: [1]
  51. Reddy, B. S.; Reddy, S. V.; Reddy, P. V.; Reddy, N. K.; Vijayalakshmi, R. P. IOP Conf. Ser.: Mater. Sci. Eng. 2015, 73, 012053. doi:10.1088/1757-899X/73/1/012053
    Return to citation in text: [1]
  52. Deka, S.; Joy, P. A. Solid State Commun. 2007, 142, 190–194. doi:10.1016/j.ssc.2007.02.017
    Return to citation in text: [1]
  53. Abrishami, M. E.; Hosseini, S. M.; Kakhki, E. A.; Kompany, A.; Ghasemifard, M. Int. J. Nanosci. 2010, 9, 19–28. doi:10.1142/S0219581X1000648X
    Return to citation in text: [1]
  54. Zhang, Y.-J.; Luo, Y.-D.; Lin, Y.-H.; Nan, C.-W. Thin Solid Films 2013, 537, 239–241. doi:10.1016/j.tsf.2013.04.082
    Return to citation in text: [1]
  55. Stefaniuk, I.; Cieniek, B.; Virt, I. Curr. Top. Biophys. 2010, 33, 221–226.
    Return to citation in text: [1]
  56. Ronning, C.; Gao, P. X.; Ding, Y.; Wang, Z. L.; Schwen, D. Appl. Phys. Lett. 2004, 84, 783–785. doi:10.1063/1.1645319
    Return to citation in text: [1]
  57. Lojkowski, W.; Leonelli, C.; Chudoba, T.; Wojnarowicz, J.; Majcher, A.; Mazurkiewicz, A. Inorganics 2014, 2, 606–619. doi:10.3390/inorganics2040606
    Return to citation in text: [1] [2]
  58. Majcher, A.; Wiejak, J.; Przybylski, J.; Chudoba, T.; Wojnarowicz, J. Int. J. Chem. React. Eng. 2013, 11, 361–368. doi:10.1515/ijcre-2012-0009
    Return to citation in text: [1]
  59. Dunne, P. W.; Munn, A. S.; Starkey, C. L.; Lester, E. H. Chem. Commun. 2015, 51, 4048–4050. doi:10.1039/C4CC10158H
    Return to citation in text: [1]
  60. Dunne, P. W.; Starkey, C. L.; Munn, A. S.; Tang, S. V. Y.; Luebben, O.; Shvets, I.; Ryder, A. G.; Casamayou-Boucau, Y.; Morrison, L.; Lester, E. H. Chem. Eng. J. 2016, 289, 433–441. doi:10.1016/j.cej.2015.12.056
    Return to citation in text: [1]
  61. Dunne, P. W.; Munn, A. S.; Starkey, C. L.; Huddle, T. A.; Lester, E. H. Philos. Trans. R. Soc. London, Ser. A 2015, 373. doi:10.1098/rsta.2015.0015
    Return to citation in text: [1]
  62. Gimeno-Fabra, M.; Hild, F.; Dunne, P. W.; Walton, K.; Grant, D. M.; Irvine, D. J.; Lester, E. H. CrystEngComm 2015, 17, 6175–6182. doi:10.1039/C5CE00628G
    Return to citation in text: [1]
  63. Sierra-Pallares, J.; Huddle, T.; Alonso, E.; Mato, F. A.; García-Serna, J.; Cocero, M. J.; Lester, E. Chem. Eng. J. 2016, 299, 373–385. doi:10.1016/j.cej.2016.04.073
    Return to citation in text: [1]
  64. Schanche, J.-S. Mol. Diversity 2003, 7, 291–298. doi:10.1023/B:MODI.0000006866.38392.f7
    Return to citation in text: [1]
  65. Byrappa, K.; Adschiri, T. Prog. Cryst. Growth Charact. Mater. 2007, 53, 117–166. doi:10.1016/j.pcrysgrow.2007.04.001
    Return to citation in text: [1]
  66. Wojnarowicz, J.; Kusnieruk, S.; Chudoba, T.; Gierlotka, S.; Lojkowski, W.; Knoff, W.; Lukasiewicz, M. I.; Witkowski, B. S.; Wolska, A.; Klepka, M. T.; Story, T.; Godlewski, M. Beilstein J. Nanotechnol. 2015, 6, 1957–1969. doi:10.3762/bjnano.6.200
    Return to citation in text: [1] [2] [3] [4]
  67. Pielaszek, R. J. Alloys Compd. 2004, 382, 128–132. doi:10.1016/j.jallcom.2004.05.040
    Return to citation in text: [1]
  68. Nanopowder XRD Processor Demo. http://science24.com/xrd/ (accessed March 10, 2016).
    Return to citation in text: [1] [2]
  69. FW1/5 4/5M method of evaluation of Grain Size Distribution by powder diffraction. http://science24.com/fw145m/ (accessed March 10, 2016).
    Return to citation in text: [1]
  70. Wejrzanowski, T.; Pielaszek, R.; Opalińska, A.; Matysiak, H.; Łojkowski, W.; Kurzydłowski, K. J. Appl. Surf. Sci. 2006, 253, 204–208. doi:10.1016/j.apsusc.2006.05.089
    Return to citation in text: [1]
  71. Xiaodi, L.; Changzhong, C.; Yiyang, Z.; Bin, J. J. Nanomater. 2013, No. 736375. doi:10.1155/2013/736375
    Return to citation in text: [1]
Other Beilstein-Institut Open Science Activities