This site uses cookies. By continuing to use this site you agree to our use of cookies. To find out more, see our Privacy and Cookies policy.
Brought to you by:
Paper

Size control mechanism of ZnO nanoparticles obtained in microwave solvothermal synthesis

, , , , and

Published 8 January 2018 © 2018 IOP Publishing Ltd
, , Citation Jacek Wojnarowicz et al 2018 Nanotechnology 29 065601 DOI 10.1088/1361-6528/aaa0ef

0957-4484/29/6/065601

Abstract

The aim of the paper is to explain the mechanism of zinc oxide (ZnO) nanoparticle (NP) size control, which enables the size control of ZnO NPs obtained in microwave solvothermal synthesis (MSS) within the size range between circa 20 and 120 nm through the control of water content in the solution of zinc acetate in ethylene glycol. Heavy water was used in the tests. The mechanism of ZnO NPs size control was explained, discussed and experimentally verified. The discovery and investigation of this mechanism was possible by tracking the fate of water molecules during the whole synthesis process. All the synthesis products were identified. It was indicated that the MSS of ZnO NPs proceeded through the formation and conversion of intermediates such as Zn5(OH)8(CH3COO)2 · xH2O. Esters and H2O were the by-products of the MSS reaction of ZnO NPs. We justified that the esterification reaction is the decisive stage that is a prerequisite of the formation of ZnO NPs. The following parameters of the obtained ZnO NPs and of the intermediate were determined: pycnometric density, specific surface area, phase purity, average particles size, particles size distribution and chemical composition. The ZnO NPs morphology and structure were determined using scanning electron microscopy.

Export citation and abstract BibTeX RIS

1. Introduction

ZnO is an n-type II–VI wurtzite semiconductor with a direct wide band gap (3.37 eV), and a large exciton binding energy (60 meV) [1]. It has unique properties and is widely used in many important industrial areas, such as varistors, gas sensors, solar cells, electrodes and catalysts [1]. The antibacterial and antifungal action of ZnO are reasons why it is an object of large interest of pharmacy and biomedicine [24]. In addition, new potential applications are sought for ZnO, e.g. in drug delivery [5], as a nanofiller (nanomodifier) of plastics [7, 8] or a modifier of lyotropic liquid crystals [9]. However, to provide ZnO with novel electrical, optical and chemical properties, the size and shape of ZnO nanostructures need to be well controlled [6]. Different fabrication methods, such as hydrothermal synthesis, precipitation process and calcination, vapour phase transport, sol-gel synthesis, pulsed laser deposition and thermal decomposition, have been widely used for obtaining ZnO nanomaterials (NMs) [10]. However, it is not easy to synthesise ZnO NMs given the heterogeneous crystallisation of ZnO crystals; the heterogeneous growth habit of ZnO crystals is seriously affected by e.g. temperature. The high temperature and other rigorous conditions have limited the nano-solid state process [11]. In contrast to the high temperature process, the 'wet organic chemistry' methods have recently aroused a wide interest, especially solvothermal synthesis methods because of an interest in developing a low-temperature, medium-scale production of ZnO NMs. Furthermore, solvothermal synthesis methods can provide various morphology and size controlled nanostructures and exert a considerable influence on the final products [1214].

One of the primary goals of the present development of nanomaterial synthesis is to obtain repeatable homogeneous nanostructures characterised by controlled properties such as size, purity, shape, morphology and content of dopant. The properties of ZnO nanostructures depend to a large extent on the method of obtaining, the conditions of synthesis, the nature of the solvent and substrates used [10]. The basis for the new approach to designing ZnO NMs and opening a field for testing of new phenomena is understanding synthesis mechanisms. As shown in the literature concerning reactions of ZnO NPs synthesis in organic solvents, there are several proposed presumed reaction mechanisms [1332].

Du et al [15] reported obtaining homogeneous spherical ZnO NPs with a controlled size ranging from 25 to 100 nm via esterification of zinc acetate and ethanol under solvothermal reaction conditions. By the several reaction conditions, Du et al [15] confirmed that by changing the reaction temperature and duration, the NPs size can be controlled. The Fourier transform infrared (FT-IR) analysis confirmed the existence of ethyl acetate (an ester) [15]. The suggested mechanism possibly contained some coupling reactions. The aforementioned authors suggested that OH ions are formed at the first stage as the products of esterification reaction of CH3COO anion with ethanol. The discussion of FT-IR results did not include the possible presence of acetic acid. An esterification reaction certainly takes place but the gathered results were insufficient to determine whether the following reaction (1) proceeds without or with an intermediate stage

Equation (1)

The above reaction model was used for explaining the control of size and shape of ZnO NPs as a result of using various organic solvents, alcohols and polyols [1618]. We argue that the esterification reaction is necessary for the formation of ZnO NPs because the released acetic acid would partly dissolve the ZnO NPs being formed. However, there is a need to take into account that the observed esterification reaction taking place in the solution of dissolved zinc acetate in ethanol can be:

  • -  
    a reaction leading to the formation of an intermediate in the ZnO synthesis;
  • -  
    a side reaction taking place as a consequence of ZnO formation;
  • -  
    a competitive reaction in relation to ZnO synthesis.

Spanhel [19] mentions identification of oxy-zinc acetate Zn4O(CH3COO)6 as an intermediate in the synthesis reaction of ZnO layers in organic solvents. The control of size and shape was explained by the change in size of the forming clusters through a change in salt concentration, synthesis conditions, nature of organic solvent, and storage conditions. According to this author [19], all the enumerated factors significantly influence the structure of ZnxOy(CH3COO)z cluster, which determines the properties of the formed ZnO NMs.

Hosono et al [20] identified another organic salt in their work. The layered hydroxy-zinc acetate (LHZA) Zn5(OH)8(CH3COO)2 · 2H2O was found to be one of the intermediates of ZnO synthesis in an organic solvent (methanol). Authors did not provide a possible mechanism of size control of the obtained nano- and microstructures.

Searson et al [2125] analysed the optical absorption spectra of ZnO NPs grown in alcohol environments. They suggest that the ZnO growth is controlled by Ostwald ripening process, i.e., small particles dissolve and redeposit onto larger particles. However, Wood et al [26] suggested the existence of a surface reaction that limited the Ostwald growth. Meulenkamp, in turn, in his papers [27, 28] concerning the examination of growth and acidic dissolution of ZnO indicates the presence of anomalous size dependencies of growth and etch rates. The abnormality observed by Meulenkamp could not be explained with the classical particle size dependent solubility predicted by the Kelvin equation. Obtaining large ZnO particle sizes cannot be explained solely by the classical dissolution-reprecipitation theory [2931]. It should be borne in mind that the ZnO particle growth process is strongly correlated with the temperature and composition of the precursor solution/suspension. Probably there are several mechanisms of ZnO growth in organic solvents, which would explain the diversity of the obtained ZnO nano- and micro-structures [10, 12].

There is only one paper where ZnO NMs growth control takes place by adding H2O [13]. The precursor in this reaction was a solution of dehydrated zinc acetate dissolved in methanol. Wang et al [13] obtained heterogeneous ZnO microstructures with foreign phases coming probably from Zn(OH)8(CH3COO)2 · 2H2O. The mechanism of ZnO crystal growth control by adding H2O was interpreted as a probable impact of water on the simultaneous hydrolysis of Zn(CH3COO)2 · 2H2O and the intermediates that formed. It was suggested that [Zn(OH)n2−n]x was a possible product of a rapid hydrolysis of Zn5(OH)8(CH3COO)2 · 2H2O. At the further stage of synthesis [Zn(OH)n2−n]x was subject to dehydration, as a result of which ZnO and H2O were formed. What is important, Wang et al [13] did not confirm the presence of Zn(OH)2 phase in the phase analysis of intermediates. The heterogeneity of the obtained crystals was explained by the varied speed of ZnO growth directions. Wang et al did not present an experimental confirmation of the suggested mechanism model in their paper [13]. It was based exclusively on XRD and scanning electron microscopy (SEM) results. The presented model of ZnO synthesis mechanism did not take into account all the obtained intermediates and by-products.

In solvothermal synthesis, the most often used precursor of ZnO NMs synthesis reaction is zinc acetate hydrate dissolved in alcohol or polyol. In general, researchers do not consider the presence of water in the precursor despite the fact that water is added to the precursor in two ways. The first way is with the hydrated zinc acetate salt, and the second way—together with the organic solvent where it occurs in trace quantities. Own research indicated a significant impact of H2O content in the precursor on the course of solvothermal reaction of ZnO NPs [12]. We observed that it was possible to control the size of the obtained ZnO NPs within the range between circa 20 and 120 nm through the change of H2O concentration in the solution of zinc acetate in ethylene glycol (EG). The research results that have been obtained and described so far in literature dealing with solvothermal synthesis reveal the lack of fundamental knowledge concerning the significance of H2O [32]. However, it seems that some amount of water is necessary for the course of reaction [33]. The main functions of water in the mechanism of ZnO solvothermal synthesis, and concerning the observed phenomenon of ZnO NP size control, have not been unequivocally explained and verified so far. It needs to be borne in mind that each mechanism of ZnO solvothermal synthesis should be considered individually. This results from the variety of types of reagents used and at the same time the quantity of obtainable intermediates and by-products [34].

The aim of our research was to explain the size control mechanism of ZnO nanoparticles obtained using the MSS method, in which the size of ZnO NPs depends on the concentration of water in the solution of zinc acetate dissolved in EG. The MSS method employed by us is quicker, purer, more energy-efficient and economical than the conventional solvothermal methods. The microwave heating applied is a contactless effective manner of feeding energy to the reaction medium located in a Teflon® reaction chamber, which results in obtaining ZnO NPs with a narrow size distribution [35]. The growing popularity of MSS is proved by the development of new types of microwave reactors [3537].

2. Experimental

2.1. Substrates

Zinc acetate dihydrate (Zn(CH3COO)2 · 2H2O, analytically pure, SKU: 112654906-1KG, Chempur); zinc acetate dihydrate (Zn(CH3COO)2 · 2H2O, analytically pure, SKU: 265490116-1KG, Avantor Performance Materials Poland S.A.); dehydrated zinc acetate (Zn(CH3COO)2, analytically pure, SKU: 112654907-250G, Chempur); ethylene glycol (EG, ethane-1,2-diol, C2H4(OH)2, pure, SKU: 114466303-5L, Chempur); diethylene glycol (DEG, 2-(2-hydroxyethoxy)ethanol, (C2H4OH)2O, pure, Chempur); triethylene glycol (TEG, 2-[2-(2-hydroxyethoxy)ethoxy]ethanol, HOCH2(C2H4O)2CH2OH, analytically pure, SKU: T59455-1L Sigma-Aldrich); tetraethylene glycol (TTEG, 2-[2-[2-(2-hydroxyethoxy)ethoxy]ethoxy]ethanol, HOCH2(C2H4O)3CH2OH, analytically pure, SKU: 110175-1KG Sigma-Aldrich) and deuterium oxide (D2O, 99.9 atom% D, SKU: 151882-100G, Sigma-Aldrich), were used. All the reagents were used without additional purification. Deionised water with specific conductance below 0.1 μS cm−1 was obtained using a deioniser (HLP 20UV, Hydrolab).

2.2. Synthesis of ZnO NPs

ZnO NPs syntheses were carried out in accordance with the procedure included in our earlier publications [12, 38, 39]. The reaction precursor, 1500 ml of solution of zinc acetate hydrate in EG (0.3037 mol dm−3), was prepared using a hot-plate magnetic stirrer (SLR, SI Analytics, Germany) at the constant temperature of 70 °C, stirring speed of 450 rpm. After complete dissolution of zinc acetate, the solution was poured to two 1000 ml PP bottles and closed tightly. After cooling down to the room temperature (RT), an analysis of water content in the precursor was carried out (1.05 ± 0.03 wt% of H2O). An appropriate calculated amount of D2O was added to the precursor to obtain the water content being 1.5 wt% and 4 wt%, and subsequently the solution was stirred again and an analysis of water content in the prepared precursor was carried out. Then 75 ml of the solution was poured into a 110 ml Teflon® reaction container and closed tightly.

The reaction, initiated with microwave radiation, was carried out in a Magnum 02-02 reactor (600 W, 2.45 GHz, ERTEC). The durations of individual synthesis reactions were 5, 6, 7.5, 10, 15, 20 and 25 min; temperature 190 °C; capacity 100%, and after that the reaction container was cooled down for 20 min. After the synthesis, the obtained powder was sedimented, rinsed three times with deionized water, centrifuged (MPW-350, MPW Med Instruments) and dried in a freeze dryer (Lyovac GT-2, SRK Systemtechnik GmbH).

3. Characterisation methods

3.1. X-ray powder diffraction

Diffraction patterns of the x-ray powder diffraction (XRD) were gathered at the RT within the range of 2 theta angle from 10° to 100° with the step of 0.02°, using the x-ray powder diffractometer (CuKα1) (X'Pert PRO, Panalytical) [41]. The full width at half maximum was determined by the Pearson VII function implemented in Fityk software, version 0.9.8. Based on the diffraction patterns, the size of crystallites was determined in the direction of the crystallographic axes a and c using Scherrer's formula [12, 39].

3.2. Crystallite size distribution

The analysis of XRD peak profile was performed using the analytical formula for polydispersive powders [42]. This technique provides four parameters: average crystallite size, deviation of the average crystallite size, dispersion of size and deviation of dispersion of sizes. Hence, a full crystallite size distribution curve and an estimation of 'thickness' of this curve (deviation bars) are obtained.

For calculating the crystallite diameter and size distribution, the Nanopowder XRD Processor Demo web application employs equations dedicated to spherical crystallites. The website http://science24.com/xrd provides an on-line tool where diffraction files can be directly dropped [43]. The files are processed on a server to extract particle size distribution for XRD peaks [44]. Unlike the standard fitting, the tool does not act in the reciprocal space but solves all sets of equations in a few auxiliary spaces simultaneously. This allows an analysis of XRD data with heavily convoluted reciprocal space peaks.

3.3. Measurement of density and specific surface area (SSA)

Skeleton density (pycnometric density) measurements were carried out using the helium pycnometer [45] (AccuPyc II 1340, FoamPyc V1.06, Micromeritics), the measurements were carried out in accordance with ISO 12154:2014 at the temperature of 25 ± 2 °C. The SSA of NPs was determined using the surface analyser (Gemini 2360, V 2.01, Micromeritics) by the nitrogen adsorption–desorption method based on the linear form of the Brunauer–Emmett–Teller isotherm equation, in accordance with ISO 9277:2010. Prior to performing measurements of density and SSA, the ZnO samples were subjected to 2 h desorption (VacPrep 061, Micromeritics), under vacuum (0.05 mbar) at the temperature of 150 °C. The samples with the intermediate were subjected to 8 h desorption under vacuum (0.05 mbar) at the temperature of 60 °C. Based on the determined SSA and skeleton density, the average size of particles defining their diameter was determined, with the assumption that all particles are spherical and identical. The equation used for calculating the average particle size can be found in our earlier publications [12, 38, 39].

3.4. Morphology characteristics

The morphology of NPs was determined using the SEM (ZEISS, ULTRA PLUS). Powder samples were coated with a thin carbon layer using the sputter coater (SCD 005/CEA 035, BAL-TEC). An internal laboratory measurement procedure was applied (P5.10, edition 6 of 26.08.2015).

The morphology of the nanopowder samples was examined using transmission electron microscopy (TEM—Talos F200X). The TEM tests using the dark field (DF) and selected area electron diffraction were conducted at 200 kV. High-resolution (HRTEM) tests were conducted at 300 kV. The specimens for the TEM observations were prepared by dropping the ethanol particle dispersion, created by an ultrasonic technique, on a carbon film supported on a 300 mesh copper grid. TEM tests were used to determine the nanoparticle size distribution. The grain size histograms were obtained by considering a region of a sample having about 200 nanocrystals and approximating the shape of each nanocrystal by a sphere. The obtained histograms were fitted to log-normal distributions [12].

3.5. Water content analysis

The quantitative analysis of water in the precursor was carried out in accordance with the assumptions of the Karl Fischer method using the coulometric titration technique with the titrator (Cou-Lo AquaMAX KF, GR Scientific) [12]. The following reagents were used for the analysis: Aquagent®Coulometric OIL, cat. no. AQ00250100, Scharlau and Aquagent®Coulometric CG, cat. no. AQ00230050, Scharlau. An internal laboratory measurement procedure was applied. Liquid samples were introduced to the titration vessel using a glass syringe (1 ml) with a Luer type needle. An analytical scales was used for weight measurement (WAA 100/C/1, RADWAG).

3.6. Gas chromatography-mass spectrometry (GC-MS)

Reaction products were identified using GC-MS (Agilent 7890 B). Before the GC, injection samples (100 μl) were dissolved in trichloromethane (1 ml). An HP-5 capillary column (5% diphenyl and 95% dimethylpolysiloxane, 30 m × 0.25 mm × 0.25 μm) was applied. The injector was operated in splitless mode. The carrier gas was He and column flow rate was 1.2 ml min−1. The following temperature programme was used for the analysis: initial temperature of 50 °C for 3 min, temperature rate of 10 °C min−1 until 240 °C and final temperature of 240 °C for 20 min.

Before the GC-MS analysis, the samples of ZnO NPs suspended matter in EG were centrifuged with the centrifugal force of 24.270 g (MPW-251, MPW Med. Instruments) for 60 min.

3.7. FT-IR spectroscopy

A Bruker Tensor 27 infrared spectrometer, equipped with an attenuated total reflectance (ATR, model: Platinium ATR-Einheit A 255), was used for the FT-IR spectroscopy analysis. For each ATR-FT-IR spectrum, 100 scans within the range of 350–4000 cm−1 with the resolution of 4 cm−1 at RT were collected.

Before the FT-IR analysis, the samples of ZnO NPs suspended matter in EG were centrifuged with the centrifugal force of 24.270 g (MPW-251, MPW Med. Instruments) for 60 min at RT.

3.8. Thermogravimetric analysis (TGA)

The TGA was carried out using the thermal analyzer (STA 449 F1 Jupiter® Netzsch) from RT to 970 °C, with a heating rate of 5 °C min−1 in an alumina crucible under synthetic air (40 ml min−1). TGA's equipment error is 0.01 mg.

3.9. Elemental analysis

The elemental analysis of C and H was carried out using the Vario EL III device by Elementar in accordance with the manufacturer's recommendations.

4. Results and discussion

4.1. Morphology

Figures 1 and 2 present representative SEM images of the ZnO NPs synthesis products. SEM images 1(a)–(c), 2(a) show the lamellar structure of the obtained intermediate. We did not observe an impact of the synthesis duration on the thickness of the lamellar structures, which was 50 ± 6 nm. When comparing synthesis products with differing contents of H2O in the precursors, differences in the obtained product types are noticeable for the same synthesis durations. For the H2O content of 1.5 wt%, lamellar structures were obtained for synthesis durations of 5, 6 and 7.5 min. For the H2O content of 4 wt%, in turn, lamellar structures were observed only in the sample with the synthesis duration of 6 min. It should be emphasised that for the synthesis duration of 5 min no solid product was obtained from the precursor with the 4% H2O content, which probably results from the impact of the 4% H2O content on the kinetics of the reaction of intermediate formation.

Figure 1.

Figure 1. SEM images of ZnO NPs synthesis products obtained from the 1.5%H2O precursor, for synthesis durations of (a) 5 min; (b) 6 min; (c) 7.5 min; (d) 10 min; (e) 15 min; (f) 20 min; (g) 25 min, respectively.

Standard image High-resolution image
Figure 2.

Figure 2. SEM images of ZnO NPs synthesis products obtained from the 4%H2O precursor, for synthesis durations of (a) 6 min; (b) 7.5 min; (c) 10 min; (d) 15 min; (e) 20 min; (f) 25 min, respectively.

Standard image High-resolution image

Figures from 1(d)–(g) and from 2(c)–(f) present homogeneous ZnO NPs with the spherical shape. Figure 1(d) shows conglomerates of ZnO NPs with the shape resembling the structure of a cauliflower. The average size of particles obtained from the 1.5%H2O precursor increases from ≈20 ± 3 nm (10 min, figure 1(d)) to ≈25 ± 5 nm (25 min, figure 1(g)). For the 4%H2O precursor, the average size of ZnO NPs increases from ≈21 ± 5 nm (7.5 min, figure 1(b)) to ≈47 ± 7 nm (25 min, figure 2(f)). When viewing SEM photos for the synthesis products from the 4%H2O precursor, we noticed growth of medium-size ZnO NPs by circa 20 nm, while for the 1.5% H2O content in the precursor—by merely circa 3–5 nm.

4.2. Phase composition

XRD tests indicated the presence of only hexagonal phase ZnO (JCPDS card No. 36-1451) in the samples with synthesis durations of 10, 15, 20, 25 min obtained from 1.5%H2O precursor (figure 2) and with synthesis durations of 7.5, 10, 15, 20, 25 min obtained from 4%H2O precursor (figure 3). Figures 3 and 4 show differences between the widths of diffraction peaks of the ZnO phase. Namely, the observed decrease of the diffraction peaks width in line with the increase in the ZnO NPs synthesis duration means at the same time an increase in the crystallite size. The phase analysis of samples with the synthesis durations of 5, 6, 7.5 min (1.5%H2O, figure 3) and with synthesis durations of 6 min (4%H2O, figure 4) indicated the existence of only the phase of the intermediate, probably of Zn5(OH)8(CH3COO)2 · 2H2O.

Figure 3.

Figure 3. XRD diffraction patterns of ZnO NPs synthesis products obtained from the 1.5%H2O precursor, for synthesis durations of 5, 6, 7.5, 10, 15, 20, 25 min and its comparison with the standard pattern of ZnO in wurtzite phase (JCPDS card No. 36-1451).

Standard image High-resolution image
Figure 4.

Figure 4. XRD diffraction patterns of ZnO NPs synthesis products obtained from the 4%H2O precursor, for synthesis durations of 6, 7.5, 10, 15, 20, 25 min and its comparison with the standard pattern of ZnO in wurtzite phase (JCPDS card No. 36-1451).

Standard image High-resolution image

The literature review revealed [4648] that the hexagonal phase of Zn5(OH)8(CH3COO)2 · 2H2O contains also peaks below 10°. In order to confirm the correctness of the phase analysis, extended XRD tests of the samples were carried out within the range of 2 theta angle from 5° to 65° at RT (figure 5). All diffraction peaks in figure 5 were attributed to the hexagonal phase of hydroxy-zinc acetate—Zn5(OH)8(CH3COO)2 · 2H2O [4648]. Zn5(OH)8(CH3COO)2 · 2H2O is a monomer of lamellar sheet compounds and therefore it is also called lamellar hydroxy-zinc acetate (LHZA) [20]. The observable shifts of the position and the changes in the peak shapes in the diffraction patterns of LHZA compounds in figure 5 result from:

  • -  
    the different quantity of OH, CH3COO function groups,
  • -  
    the quantity of crystallisation water 'x', which may range from 1.5 to 4, depending on the synthesis method employed [20, 4751],
  • -  
    obtaining a hybrid material, e.g. Zn5(OH)8(CH3COO)2(C2H4(OH)2)2 · 2H2O, through the process of EG intercalation between the layers of LHZA [52].

Figure 5.

Figure 5. Comparison of XRD diffraction patterns of the intermediates of the ZnO NPs synthesis and its comparison with the standard pattern of Zn5(OH)7.9(CH3COO)2.1 · 1.55H2O (JCPDS card No. 56-0569).

Standard image High-resolution image

In addition, possible changes in the lattice parameters in LHZA caused by the exchange of OH or CH3COO groups by EG should be taken into account. Such cases are known where EG created a covalent bond with a metal in metal hydroxide layers, e.g. in brucite, boehmite and layered double hydroxides [5358].

The interlamellar distances of intermediates that we calculated and presented in table 1 increase in line with the increase in the synthesis duration and with the increase in the H2O content in the precursor. The interlamellar distance (001) of all the obtained samples are lower than 14.72 Å, i.e. for the interlamellar distance (001) of the reference sample of LHZA (JCPDS card No. 56-0569). Kasai et al [52] report that the decrease of the interlamellar distance (001) value below 14.72 Å in LHZA, which they observed, can be caused by a reduction of the quantity of CH3COO groups and of crystallisation H2O. The increase in the interlamellar distance (001) above 14.72 Å, in turn, is explained with EG intercalation [52]. We believe that, based on the results of interlamellar distance (001), intercalation of EG between the layers of LHZA cannot be excluded even if the interlamellar distance (001) value is below 14.72 Å without verifying the lack of presence of EG in the obtained LHZA samples. XRD results indicate that the LHZA compounds obtained by us were characterised by differing stoichiometry (different quantity of CH3COO, OH, H2O groups). The interlamellar distance (001), and at the same time stoichiometry, that was most similar to the reference sample of LHZA (JCPDS card No. 56-0569) was exhibited by ZnO-4%H2O-6 min sample (table 1).

Table 1.  Interlamellar distance in the obtained intermediates (LZHA).

Sample Interlamellar distance, dhkl (Å)
  001 002 003
ZnO-1.5%H2O-5 min 12.87 6.53 4.39
ZnO-1.5%H2O-6 min 13.41 6.72 4.53
ZnO-1.5%H2O-7.5 min 13.80 6.91 4.69
ZnO-4%H2O-6 min 14.06 7.04 4.75
Reference sample Zn5(OH)7.9(CH3COO)2.1 · 1.55H2O JCPDS card No. 56-0569 14.68 7.39 4.98

4.3. Chemical analysis and empirical formula

The empirical formulas of compounds were determined based on the obtained mass contents of C, H, Zn, O. The elemental analysis of C and H was carried out by the combustion method. Zn contents were calculated based on the results of mass loss of LHZA samples after their thermal decomposition to ZnO at 950 °C (figure 6). Contents of O, in turn, were calculated based on the following formula: O% = 100% − C% − H% − Zn% (table 2). We assumed that C% content in LHZA corresponds only to the carbon content in CH3COO groups.

Figure 6.

Figure 6. TGA profiles for LHZA samples.

Standard image High-resolution image

Table 2.  Chemical composition of LHZA samples.

  C (wt%) H (wt%) Zn (wt%) O (wt%)   Mass loss (wt%)
Sample Results Calca Results Calca Results Calca Results Calca Empirical formula Results TGA Calca
ZnO-1.5%H2O- 5 min 5.67 5.70 3.01 3.32 51.75 51.73 39.57 39.25 Zn5(OH)8.50 (Ac)1.50 · 4H2O 35.59 35.62
ZnO-1.5%H2O- 6 min 10.76 10.63 2.97 3.02 50.08 49.93 36.19 36.41 Zn5(OH)7.10(Ac)2.90 · 2H2O 37.67 37.85
ZnO-1.5%H2O-7.5 min 10.88 10.88 2.94 3.02 49.39 49.74 36.79 36.35 Zn5(OH)7.02 (Ac)2.98 · 1.95H2O 38.52 38.09
ZnO-4%H2O-6 min 6.98 6.95 3.11 3.22 51.35 51.45 38.56 38.38 Zn5(OH)8.16(Ac)1.84 · 3.4H2O 36.09 35.97
LHZA reference 7.78 2.92 52.99 36.31 Zn5(OH)8(Ac)2 · 2H2O 34.05

aCalculated.

In order to verify the determined empirical formulas of LHZA samples, theoretical calculations of C, H, Zn, O contents and of mass loss after their decomposition to ZnO were performed (table 2). Calculated percentages in table 2 are based on the ideal formulas. The obtained results of total %C content in the samples are consistent with the calculations that verified them (table 2). However, in the case of samples with C content exceeding 7.78%, i.e. for ZnO-1.5%H2O-6 min and ZnO-1.5%H2O-7.5 min, we presumed that the excess %C content was caused by the presence of EG in these samples. This was confirmed by TGA results (figure 6), where a shift of thermal stability towards higher temperatures was observed for ZnO-1.5%H2O-6 min and ZnO-1.5%H2O-7.5 min samples, which results from a high boiling point of EG (197.3 °C) [52]. EG in these samples may have been physically adsorbed on LHZA surface or intercalated between the layers of LHZA [46, 58]. The total thermal decomposition of the stoichiometric LHZA compound leads to a theoretical mass loss equal to 34.05% (table 2). Total mass losses observed in our samples ranged from 35.59% to 38.52% (table 2). Results of works by different research groups [20, 4649, 51, 59], in turn, concerning the value of total LHZA mass loss were as follows: 32.15%, 37.4%, 38.4%, 39.7% and 43.9%. This means that the majority of the reported syntheses led to obtaining non-stoichiometric LHZA. Hosono et al [20] report that stoichiometry of LHZA depends on synthesis conditions. Kasai et al [52] note, in turn, that stoichiometry of LHZA changes depending on the time of its storage in EG at RT. Moreover, Moezii et al [49] observed partial decomposition of LHZA to ZnO during its storage in ethanol. The presented results prove that the stoichiometry of LHZA compounds depends not only on synthesis parameters but also on the duration of sample storage directly after the synthesis in the form of suspended matter in alcohols and polyols.

The results of the elemental analysis of our samples are consistent with the results of the phase analysis, where the shifts of LHZA phase peaks were explained with obtaining different stoichiometry (figure 5). The change in stoichiometry can be explained with a different share of the nucleation stage and the growth stage on the moment of synthesis finish for individual LHZA samples.

4.4. FT-IR tests and GC-MS analysis

FT-IR spectra and spectral assignments of the samples are presented in figures 7, 8 and in table 3, respectively. The characteristic band coming from the stretching vibrations of –OH is observed primarily within the range between 2900 and 3600 cm−1. Depending on the quantity of –OH groups, the process of intermolecular hydrogen bond formation intensifies, which can be an explanation of the shift of a broad band of -OH group from 3380 to 3450 cm−1 in the obtained results (figure 7). Two bands with the maximums at 2930 and 2867 cm−1 coming from symmetric and asymmetric stretching vibrations of C–H prove the presence of C2H4(OH)2 or CH3COO- groups. Bands coming from the stretching vibrations of C–H were visible only in two LZHA samples with the carbon content exceeding 7.78%, which proved the presence of C2H4(OH)2 in these samples. The spectra of ZnO-1.5%H2O-6 min, ZnO-1.5%H2O-7.5 min and ZnO-4%H2O-6 min in figure 7 show a band at 2360 cm−1, which comes from asymmetric stretching vibrations in CO2, which means that a change in CO2 content in atmospheric air occurred during the measurements of the samples (measurement background). Strong absorption bands with wavenumbers of 1573 and 1404 cm−1 come from asymmetric and symmetric stretching vibrations of C=O group in CH3COO. Bands with frequencies of 1324 cm−1 were attributed to symmetrically bending vibrations of –CH3 groups in CH3COO. The band within the range of 1083–1030 cm−1 observed in the spectra of samples with LHZA phase comes from asymmetrical stretching vibrations of C–O groups in C2H4(OH)2. Presence of C2H4(OH)2 in the samples is confirmed also by changes of band intensity within the range of 1083–1030 cm−1, which are in proportion to the changes in carbon content (table 2, figure 7). FT-IR spectra of LHZA compounds obtained by us are consistent with the results of other authors [20, 46, 5158] and confirmed the presence of CH3COO, OH groups and EG.

Figure 7.

Figure 7. FT-IR spectra at RT of dry samples of LHZA and EG, H2O, D2O.

Standard image High-resolution image
Figure 8.

Figure 8. FT-IR spectra at RT of solution of the 1.5%H2O precursor (synthesis duration 0 min) and of solution obtained after centrifuging the post-reaction suspended matter of 1.5%H2O (synthesis duration 25 min).

Standard image High-resolution image

Table 3.  FT-IR spectral assignments for samples of LZHA, D2O, H2O, EG.

Wavenumbers (cm−1) Assignment
3450−3380 O–H-stretching vibrations
2453 O–D-stretching vibrations
2930 C–H-asymmetric stretching vibrations
2867 C–H-symmetric stretching vibrations
2360 O=C=O-asymmetric stretching vibrations
1724 C=O-stretching vibrations
1627−1596 O–H-bending vibrations
1573 COO-asymmetric stretching vibrations
1404 COO-symmetric stretching vibrations
1324 CH3-symmetric bending vibrations
1220−1200 O–D-bending vibrations
1083 and 1033 C–O-asymmetric stretching vibrations

The results of FT-IR, TGA tests and of the elemental analysis proved unambiguously that EG was present in two obtained LHZA samples. Nevertheless, it is possible that trace quantities of EG are present in the remaining samples. The empirical formulas of LHZA compounds that we determined can be flawed due to the failure to take into account EG content in their composition (table 2). EG in the samples could have been physically adsorbed on LHZA surface or intercalated between the layers of LHZA. All four LHZA samples were subjected to the same preparation process of both rinsing and drying, and therefore we believe that there was a greater likelihood that the EG present in two LHZA samples could be present within the interlamellar distance (001) rather than on LHZA surface.

The lack of absorption bands coming from vibrations of D2O in dry LHZA samples (figure 7) turned out to be a breakthrough result of FT-IR analyses. We do not observe absorption bands coming from D2O vibrations even for a sample obtained from a precursor in which the total H2O content was 4%H2O, where D2O accounted for 3% of that content. This may mean that water added (D2O) to the precursor solution did not participate in the chemical reaction of LHZA synthesis. Therefore, –OH groups (figure 7) present in LHZA may come exclusively from the reaction of Zn(CH3COO)2 · 2H2O with EG. In FT-IR spectra, in turn, we do not observe absorption bands coming from the bending vibrations of crystallisation H2O or D2O within the range between 1627 cm−1 and 1596 cm−1 or between 1220 cm−1 and 1200 cm−1, respectively. This may mean that absorption bands of crystallisation water (H2O or D2O) in LHZA are not visible in the obtained spectra. Our presumptions were confirmed during the tests of a sample of the 1.5%H2O precursor with 0.45% D2O content (figure 8), where absorption bands coming from D2O vibrations were virtually unnoticeable. Therefore, we presume that the quantity of crystallisation water (D2O or H2O) in the tested powder samples was too low for the detection threshold of FT-IR spectroscopy.

Figures 8 and 9 show FT-IR spectra of solutions of ZnO NPs precursor (0 min) and of solutions obtained by centrifuging the post-reaction suspended matter (25 min). When comparing both spectra, it is evident that such a product is formed in the ZnO NPs synthesis that had the absorption band with the wavenumber of 1724 cm−1. Three groups of chemical compounds can be the potential source of absorption characterised by such vibration frequencies. One of the most probable by-products of ZnO NPs synthesis is an ester, where the range of the absorption band of the stretching vibrations of the double bond of carbon with oxygen (C=O) from saturated esters is 1715–1750 cm−1. Another possible by-product of the synthesis is acetic acid, with the range of the absorption band of the stretching vibrations of C=O being 1680–1750 cm−1. The least probable reaction product is an aldehyde, where carbonyl groups of aliphatic aldehydes indicate absorption at about 1720–1740 cm−1.

Figure 9.

Figure 9. FT-IR spectra at RT of solution of the 4%H2O precursor (synthesis duration 0 min) and of solution obtained after centrifuging the post-reaction suspended matter of 4%H2O (synthesis duration 25 min).

Standard image High-resolution image

In order to determine the by-product of the ZnO NPs synthesis, solution samples were subject to GC-MS analysis. Figure 10(a) presents a chromatogram of the ZnO-1.5%H2O-25 min sample (solution), where two peaks with retention times of 4.8 and 7.5 min are visible. Figure 10(b) contains mass spectra. Two peaks with different retention times in the obtained chromatogram mean obtaining two liquid products in the solvothermal synthesis of ZnO. The first peak with retention time of 4.8 min was identified as an ester, ethanediol monoacetate (HOC2H4OOCH3), while the second one with retention time of 7.5 min is another ester, ethanediol diacetate (C2H4(OOCH3)2). Analogous results were obtained for the ZnO-4%H2O-25 min sample (solution). The GC-MS analysis indicated unequivocally that esters are formed during the ZnO NPs synthesis. The presence of esters as a by-product in the solvothermal reaction of zinc acetate in various alcohols was also detected by Tonto et al [16].

Figure 10.

Figure 10. (a) Chromatogram for the solution obtained after centrifuging the post-reaction suspended matter of ZnO-1.5%H2O-25 min. (b) and (c) the GC-MS spectrum of the solution of ZnO-1.5%H2O-25 min. Sample (d) and (e) matches to the GC-MS spectrum results.

Standard image High-resolution image

4.5. Density, SSA and average size and size distribution of NPs

The skeleton density of LHZA samples obtained from the 1.5%H2O precursor ranged from 2.43 to 2.57 g cm−3 for synthesis durations from 5 to 7.5 min (table 4). The density of the LHZA sample obtained from the 4%H2O precursor for the synthesis duration of 6 min was 2.87 g cm−3. The SSA value of LHZA samples from the 1.5%H2O precursor ranged from ≈59 to ≈40 m2 g−1 for synthesis durations from 5 to 7.5 min. For the sample obtained from the 4%H2O precursor, the SSA value was ≈48 m2 g−1. Reinoso et al [61], in turn, obtained LZHA with the SSA of 25 m2 g−1 by the titration method, which once again reveals a considerable impact of the synthesis method on the properties of the obtained LZHA. LZHA structures were lamellar shaped (x, y, z), which made it difficult to calculate average sizes from SSA and from methods based on XRD results, where the spherical shape of the obtained NPs is required. The differences in the results of density and SSA between the obtained LZHA samples result probably from several factors:

  • -  
    different stoichiometry (table 2);
  • -  
    possible presence of an amorphous phase;
  • -  
    different porosity;
  • -  
    presence of substances that were not evaporated at the applied desorption temperature (60 °C, 8 h, vacuum);
  • -  
    partial decomposition of LZHA resulting from sample desorption (60 °C, 8 h, vacuum).

Table 4.  Characteristics of synthesis products of ZnO NPs samples.

Precursor
  Specific surface area, as (m2 g−1) Skeleton density, ρs ± σ (g cm−3) Average particles size from SSA BET, d ± σ (nm) Average crystallites size, Scherrer's formula, da, dc (nm) Average crystallites size, Nanopowder XRD Processor Demo, d ± σ (nm)
Synthesis duration (min) 1.5%H2O 4%H2O 1.5%H2O 4%H2O 1.5%H2O 4%H2O 1.5%H2O 4%H2O 1.5%H2O 4%H2O
5 59.4 2.43 ± 0.04
6 39.6 48.5 2.57 ± 0.04 2.87 ± 0.05
7.5 42.5 74.7 2.53 ± 0.03 5.12 ± 0.03 16 ± 1 16a, 20c 18 ± 5
10 95.3 44.9 4.98 ± 0.04 5.32 ± 0.03 13 ± 1 25 ± 1 13a, 15c 26a, 37c 17 ± 7 29 ± 9
15 75.3 28.9 5.16 ± 0.04 5.40 ± 0.02 15 ± 1 29 ± 1 16a, 20c 34a, 51c 18 ± 5 38 ± 13
20 52.2 26.5 5.22 ± 0.03 5.42 ± 0.04 22 ± 1 43 ± 1 20a, 23c 35a, 51c 21 ± 6 39 ± 12
25 51.9 25.9 5.29 ± 0.03 5.42 ± 0.03 22 ± 1 43 ± 1 21a, 26c 35a, 53c 23 ± 6 41 ± 14

The calculated theoretical density of ZnO is 5.61 g cm−3. The skeleton density of ZnO NPs samples obtained from the 1.5%H2O precursor ranged from 4.98 to 5.29 g cm−3 for the range of NPs sizes from 13 to 22 nm (table 4). For samples of ZnO NPs synthesised from the 4%H2O precursor, the density ranged from 5.12 to 5.42 g cm−3 for the range of NPs sizes from 16 to 43 nm. The visible correlation of density and size of NPs is known and was discussed in detail in our previous publications [12, 60].

The results of average size are summarised in table 4. The average size of ZnO NPs obtained from the 1.5%H2O precursor calculated based on SSA and density results increased from 13 to 22 nm in line with the reaction duration progress from 10 to 25 min. The average crystallite size for the same samples calculated by the Scherrer method ranged from 13–15 to 21–26 nm, while using the Nanopowder XRD Processor Demo web application it ranged from 17 ± 7 nm to 23 ± 6 nm. Figure 11 presents crystallite size distributions of the obtained ZnO samples. Figure 12(a) displays the size distribution of ZnO NPs for synthesis duration of 25 min obtained based on TEM tests by the DF technique. The average size of ZnO NPs determined for that sample by TEM analysis was 20 ± 2 nm and coincided with the values obtained by other methods.

Figure 11.

Figure 11. Crystallite size distribution, obtained using Nanopowder XRD Processor Demo, pre⋅α⋅ver.0.0.8, © Pielaszek Research: (a) 1.5% H2O precursor, (b) 4% H2O precursor.

Standard image High-resolution image
Figure 12.

Figure 12. The histogram of the particle size distribution of ZnO (TEM method): (a) 1.5% H2O 25 min, (b) 4% H2O 25 min.

Standard image High-resolution image

The increase in the average size of ZnO particles/crystallites obtained from the 4%H2O precursor for the reaction duration progress from 7.5 to 25 min was as follows:

  • -  
    from 16 to 43 nm, calculation from SSA and density results;
  • -  
    from 16–20 to 35–53 nm, Scherrer method;
  • -  
    from 18 ± 5 to 41 ± 14 nm, Nanopowder XRD Processor Demo web application;
  • -  
    40 ± 1 nm for synthesis duration of 25 min, DF technique (TEM).

The average sizes calculated from SSA and density results are most representative in terms of the quantity of the tested sample. When comparing two methods of XRD result conversion, Scherrer's formula and Nanopowder XRD Processor Demo, the obtained results fit in the standard deviation of those methods. The results of average NPs sizes obtained by converting the SSA and density are consistent with the accuracy of several nanometres with the results of the methods based on the XRD analysis. Average particle size and size distributions obtained by the TEM method for ZnO-1.5%H2O-25 min and ZnO-4%H2O-25 min samples fit in the standard deviations of all the used methods (table 4, figures 11 and 12). A conclusion may be drawn from the obtained results that crystallite sizes were similar to particle sizes, which means that the obtained ZnO NPs are single crystals.

4.6. Quantity balance of products

The following products of zinc acetate reaction with EG were identified: the primary product—ZnO NPs, esters (HOC2H4OOCH3 and C2H4(OOCH3)2) being by-products, and the intermediate—Zn(OH)8(CH3COO)2 · xH2O. Another expected by-product of the synthesis was H2O. It is one of the products of the esterification reaction, i.e. the reaction of an acid with an alcohol or a polyol. The formation of H2O as a by-product of ZnO synthesis was confirmed by an analysis in the Karl Fischer titrator. Table 5 contains the results of quantitative analyses of the products (H2O, ZnO) created as a result of the synthesis. After 25 min of the synthesis reaction, the obtained ratio of the formed moles nH2O/nZnO in the samples was ≈2. In accordance with the internal laboratory measurement procedure, all obtained results of H2O content determination require a correctness confirmation. For this purpose, we verified the possible side reactions of the formed synthesis products with the applied reagents of coulometric titration. It turned out that for an ester sample (ethyl acetate, analytically pure, Chempur) and for (CH3COO)2Zn · 2H2O sample the quantity of the determined H2O was consistent with the quantity declared by the producers (table 6). However, for desorbed dry samples of ZnO NPs, where the actual values was ≈0% of H2O, the result of the H2O content analysis revealed between 22.6% and 22.9%. The obtained result is consistent with the value of molar mass ratio of MH2O/MZnO·100%, being 22.1%. This meant that during the measurements of H2O content in post-reaction suspended matter, also ZnO participated in the chemical reaction with the reagents apart from H2O. The titrator converted the ZnO quantity which was present in the tested samples as an additional H2O content. After allowing for the chemical reaction of ZnO with the titrator's reagents, the actual ratio of the mole quantities being formed nH2O/nZnO was 1. In summary: in the synthesis reaction, 1 mole of zinc acetate in the synthesis reaction gives 1 mole of ZnO, 1 mole of H2O and circa 2 moles of esters, when assuming the formation of HOC2H4OOCH3 and small quantities of C2H4(OOCH3)2. The general equation (2) of the microwave solvothermal synthesis reaction of ZnO NPs in EG is as follows:

Equation (2)

Table 5.  Quantity balance of the ZnO NPs synthesis reaction.

Sample ZnO mass obtained in synthesis (g) Theoretical ZnO mass 100% efficiency (g) ZnO NPs synthesis efficiency (%) Quantity of formed H2O as a sum of nH2O and nZnO (mol) Actual quantity of formed nH2O (mol) Quantity of formed nZnO (mol) nH2O/nZnO ratio
ZnO-1.5%H2O-25 min 1.7174 1.7530 98.0 0.0445 0.0223 0.0211 1.05
ZnO-4%H2O-25 min 1.6263 1.7318 93.9 0.0401 0.0201 0.0200 1.01

Table 6.  Verification of results of analyses of H2O content in samples with the use of Karl Fischer titrator.

Sample Result of H2O content analysis (wt%) H2O content consistent with the actual composition (wt%) Ratio of molar masses H2O/ZnO 100%
Ethyl, acetate (CH3COOC2H5, 100%) 0.01 ± 0.01 0.1 22.14
(CH3COO)2Zn · 2H2O 16.9 ± 0.3 16.4  
ZnO-1,5%H2O-25 min, dry powder 22.7 ± 0.2 0  
ZnO-4%H2O-25 min, dry powder 22.8 ± 0.2 0  

It needs to be emphasised that EG fulfilled a double function in the synthesis: the zinc acetate solvent and the substrate, and its quantity was 58 times as much as that of zinc acetate. The efficiency range of ZnO NPs synthesis is 94%–98% (table 5). The observed differences in the efficiency of ZnO NPs synthesis reaction for different H2O contents in the precursor may be caused by several factors:

  • -  
    complex preparation process of samples;
  • -  
    different H2O contents in ZnO NPs samples during their weighing.

4.7. Size control mechanism

The mechanism of size control of ZnO NPs obtained by the microwave solvothermal synthesis may be divided into four stages:

  • (1)  
    dissolution of zinc acetate in EG (3), (4), preparation of the precursor with a specified H2O content (5)–(7):
    Equation (3)
    Equation (4)
    Equation (5)
    Equation (6)
    Equation (7)
  • (2)  
    formation (8)–(11) and growth of the intermediate (12):
    Equation (8)
    or possibly e.g.
    Equation (9)
    Equation (10)
    Equation (11)
    Equation (12)
    nH2O comes from the simultaneous esterification reaction (13) or (14)
    Equation (13)
    Equation (14)
  • (3)  
    achievement of equilibrium constant of the ester hydrolysis reaction for equation (15) and at the same time of equilibrium constant of the esterification reaction (13) and decomposition of the intermediate caused by temperature (16)
    Equation (15)
    Equation (16)
  • (4)  
    growth of existing ZnO NPs (17), (18)
    Equation (17)
    Equation (18)
    The general equation of the microwave solvothermal synthesis reaction of ZnO NPs in EG that takes into account obtaining only HOC2H4OOCH3 ester, is as follows:
    Equation (19)
    The above mechanism of ZnO NPs synthesis, which is illustrated in figure 13, can be explained as follows:
    • 1.  
      The reaction precursor is a solution obtained by dissolving (CH3COO)2Zn · 2H2O in EG at an increased temperature. The dissolution of (CH3COO)2Zn · 2H2O in EG may not be treated solely as a physical process since during the dissolution of (CH3COO)2Zn · 2H2O in EG probably (CH3COO)2Zn · 2C2H4(OH)2 is formed [63, 64] as a result of the reaction of (CH3COO)2Zn · 2H2O with EG, during which EG replaces H2O without disturbing the zinc-acetate bond [63, 64]. H2O present in the precursor, in turn, causes dissociation (5) of (CH3COO)2Zn, as a result of which CH3COOH and (CH3COO)(OH)Zn are formed in the subsequent hydrolysis reactions (6), (7).
    • 2.  
      The intermediate, Zn5(OH)8(CH3COO)2 · xH2O, is formed as a result of the course of simultaneous reactions (8)–(11) and (13), (14), where an ester, (HOC2H4OOCH3), is one of the by-products of all these reactions. FT-IR tests did not reveal the presence of OD group in the intermediate, which makes us conclude that the water present in the precursor did not participate in the reaction of formation of OH groups in Zn5(OH)8(CH3COO)2 · xH2O. The 'Zn5(OH)8(CH3COO)2' part in Zn5(OH)8(CH3COO)2 · xH2O is formed, thus, as a result of the reactions (8, 9, 10, 11) of zinc acetate or its derivatives with EG. 'Crystallisation H2O' in Zn5(OH)8(CH3COO)2 · xH2O, in turn, is formed as a result of the simultaneous reaction (13), (14) of acetic acid with EG or acetic acid with (CH3COO)2Zn·2C2H4(OH)2. Zn5(OH)8(CH3COO)2 · xH2O compound is a heterogeneous catalyst of the esterification reactions [61, 62]. It should be stressed that H2O formed from the esterification reaction (13), (14), which took place on the surface of Zn5(OH)8(CH3COO)2 · xH2O, was at the same time the substrate of reactions (8)–(11) of formation of Zn5(OH)8(CH3COO)2 · xH2O compound. Equation (12) presents an example course of growth of the intermediate, where the formation and subsequent growth of Zn5(OH)8(CH3COO)2 · xH2O occurred thanks to the course of the esterification reaction.
    • 3.  
      Concentrations of the reagents HOC2H4OOCH3, CH3COOH, H2O and EG are correlated by means of the equilibrium constant. The equilibrium constant describes only equilibrium states of reversible chemical reactions. The equation of the equilibrium constant of the hydrolysis reaction (15) of the formed esters is as follows: KeqH =$\tfrac{[{{\rm{C}}}_{2}{{\rm{H}}}_{4}{({\rm{OH}})}_{2}]\cdot [{{\rm{CH}}}_{3}{\rm{COOH}}]}{[{{\rm{HOC}}}_{2}{{\rm{H}}}_{4}{{\rm{OOCH}}}_{3}]\cdot [{{\rm{H}}}_{2}{\rm{O}}]}.$ The equation of the equilibrium constant of the esterification reaction (13) is defined as follows: KeqE =$\tfrac{[{{\rm{HOC}}}_{2}{{\rm{H}}}_{4}{{\rm{OOCH}}}_{3}]\cdot [{{\rm{H}}}_{2}{\rm{O}}]}{[{{\rm{C}}}_{2}{{\rm{H}}}_{4}{({\rm{OH}})}_{2}]\cdot [{{\rm{CH}}}_{3}{\rm{COOH}}]}.$ The presented equilibrium constants are closely correlated, KeqE = $\tfrac{1}{{{\rm{K}}}_{{\rm{eqH}}}}.$ Only chemically unbound H2O takes part in the reaction of ester hydrolysis. Therefore, we can introduce water concentration to the derived equations of the equilibrium constants of the reactions, which concentration decreases in line with the progress of the intermediate formation/growth reaction. The quantity of the unreacted H2O in the precursor determines the moment when the equilibrium value KeqH and KeqE is achieved. The reactions (8)–(11) of creation or growth (12) of Zn5(OH)8(CH3COO)2 · xH2O occurring simultaneously with the esterification reaction are not equilibrium reactions, so they were still occurring on the moment of achievement of KeqE equilibrium value. The further progress of the reactions (8)–(11) of formation and growth (12) of Zn5(OH)8(CH3COO)2 · xH2O without the formation of H2O in the esterification reactions (10), (11) led to rapid decomposition of Zn5(OH)8(CH3COO)2 · xH2O (16) to nuclei of ZnO crystals (NPs), H2O and esters.
    • 4.  
      The formed homogeneous spherical nuclei of ZnO crystals (NPs) sized 16 ± 2 nm grew until the unreacted substrates (zinc acetate or its derivatives) were exhausted, which is shown in equations (17), (18) and figure 14. By-products of the growth of ZnO (NPs) crystals are H2O and esters. The formed ZnO is an autocatalyst of the esterification reaction and at the same time of the growth of ZnO NPs in the described mechanism [65], which explains such a high efficiency of (94%–98%) microwave solvothermal synthesis.

Figure 13.

Figure 13. A reference drawing presenting the mechanism of microwave solvothermal synthesis of ZnO NPs.

Standard image High-resolution image

The decomposition of Zn5(OH)8(CH3COO)2 · xH2O to nuclei of ZnO (NPs) crystals sized 16 ± 2 nm was confirmed by numerous experiments. However, the mechanism of formation of so homogeneous nuclei of ZnO (NPs) crystals with repeatable sizes through decomposition of Zn5(OH)8(CH3COO)2 · xH2O is unknown to us. This requires further research.

The ZnO NPs size control in MSS arising from the change of H2O content in the precursor is a result from the shift of the quantitative equilibrium of esterification reaction products (KeqE). The change of the water quantity in the precursor is inversely proportional to the quantity of the obtained Zn5(OH)8(CH3COO)2 · xH2O (figure 14). The following conclusion emerges: the least the H2O content in the precursor, the greater quantity of the intermediate is obtained on the moment of achievement of the equilibrium state KeqE (figure 14). At the same time the quantity of unreacted zinc acetate, which causes the growth of ZnO NPs formed as a result of decomposition of the intermediate, is lower.

Figure 14.

Figure 14. A reference phase diagram for the course of the ZnO NPs synthesis reaction, taking into account H2O content loss caused by zinc acetate hydrolysis, temperature ≈190 °C (measurement under the cup bottom in the Ertec reactor, model 02-02 [12]).

Standard image High-resolution image

Water in the reaction of (CH3COO)2Zn with EG acts as a homogeneous catalyst. Its action consists in a reaction with the substrates, as a consequence of which a temporary intermediate, Zn5(OH)8(CH3COO)2 · xH2O, is formed, which is unstable, thanks to which it reacts further with the formation of ZnO and reconstruction of H2O (catalyst).

The quantity of the necessary catalyst ${n}_{{\rm{REH}}2{\rm{O}}}$ to obtain Zn5(OH)8(CH3COO)2 · 2H2O, assuming 100% reaction efficiency, can be calculated based on equation (8), i.e. ${n}_{{\rm{REH}}2{\rm{O}}}=\tfrac{{{n}}_{({\rm{CH}}3{\rm{COO}})2{\rm{Zn}}}}{5}\cdot 2.$ However, theoretically the possibility should be considered when the H2O quantity in the precursor is lower than the necessary quantity of ${n}_{{\rm{REH}}2{\rm{O}}}.$ This is possible only when dehydrated zinc acetate and dehydrated EG are used for the synthesis. If such a case takes place, then the change of H2O quantity in the precursor will be proportional to the change of the quantity of the obtained Zn5(OH)8(CH3COO)2 · xH2O (figure 14). This results from the H2O loss caused by the zinc acetate hydrolysis.

4.8. Verification of the size control mechanism

The verification of the mechanism concerned the following aspects:

(1) According to the mechanism we propose, with minimum H2O quantities in the precursor we observe a negligible increase in ZnO NPs size. It should be remembered, though, that the ZnO synthesis reaction cannot occur in the precursor without trace quantities of H2O due to the fact that H2O is the catalyst of the ZnO synthesis. The mechanism was verified experimentally by obtaining a precursor with the H2O content of 0.065%. This precursor was prepared by dissolving dehydrated zinc acetate in EG in a tightly closed conical flask. The ZnO NPs synthesis lasted 25 min at the temperature of 190 °C. The expected average size of ZnO NPs from the completed synthesis was 17 ± 7 nm, and the value of the SSA of NPs—circa 95 m2 g−1. ZnO NPs formed on the moment of decomposition of the intermediate (table 7) are characterised by such properties. The average size of the obtained particle determined by four methods for the ZnO NPs sample in the experiment was 17 ± 5 nm and was consistent with our presumptions arising from the presented mechanism. Figures 15(a) and (b) show a narrow ZnO NPs size distribution, which ranges from 5 to 40 nm. However, the SSA of the obtained ZnO NPs was 74 m2 g−1 and was lower by 21 m2 g−1 than the expected surface area (table 7). This meant that the particle size increased during the synthesis, probably caused by the too small H2O quantity in comparison to the theoretical quantity necessary for obtaining stoichiometric LHZA with the efficiency of 100% (table 8). Table 8 contains quantities of moles of (CH3COO)2Zn and H2O which were present in the precursor, and quantities that were theoretically required for obtaining only Zn5(OH)8(CH3COO)2 · 2H2O with 100% reacted substrates.

Table 7.  Characteristics of synthesis products of NPs ZnO-0.065%H2O-25 min and ZnO-1.5%H2O-10 min.

Sample Specific surface area, as (m2 g−1) Skeleton density, ρs ± σ (g cm−3) Average particles size from SSA BET, d ± σ (nm) Average crystallites size, Scherrer's formula, da, dc (nm) Average crystallites size, Nanopowder XRD Processor Demo, d ± σ (nm) Average particles size from TEM, d ± σ (nm)
ZnO-0.065%H2O-25 min 74.0 5.04 ± 0.03 16 ± 1 15a, 20c 17 ± 5 13 ± 1
ZnO-1.5%H2O-10 min (smallest obtained ZnO NPs) 95.3 4.98 ± 0.03 13 ± 1 13a, 15c 17 ± 7
Figure 15.

Figure 15. (a) ZnO-0.065%H2O-25 min crystallite size distribution, obtained with the Nanopowder XRD Processor Demo, (b) the histogram of the particle size distribution of ZnO0.065%H2O-25 min (TEM method).

Standard image High-resolution image

Table 8.  Quantity balance.

Sample Quantity of (CH3COO)2Zn in the precursor (mol) H2O quantity in the precursor (mol) Required H2O quantity for obtaining Zn5(OH)8(CH3COO)2 · 2H2O with the 100% efficiency (mol)
ZnO-0.065%H2O-25 min 0.0228 0.0032 0.0092

(2) In accordance with the presented mechanism, the change of the zinc acetate concentration in the precursor with the constant H2O content may not considerably contribute to a change of the average size of the obtained ZnO NPs. Based on the equilibrium constant KeqE = $\displaystyle \frac{[{{\rm{HOC}}}_{2}{{\rm{H}}}_{4}{{\rm{OOCH}}}_{3}]\cdot [{{\rm{H}}}_{2}{\rm{O}}]}{[{{\rm{C}}}_{2}{{\rm{H}}}_{4}{({\rm{OH}})}_{2}]\cdot [{{\rm{CH}}}_{3}{\rm{COOH}}]}$ it may be stated that an increase in the quantity of zinc acetate with the constant H2O content will cause only a shift of the equilibrium towards a greater quantity of products. This results from the 58 times greater quantity of EG relative to zinc acetate. This means that an increase in the zinc acetate concentration will cause a proportional decrease of H2O and EG quantity in the precursor, with a proportional increase in the quantity of CH3COOH and HOC2H4OOCH3, and at the same time the quantity of the intermediate. The following correlation should be fulfilled for the quantity of the reacted zinc acetate:

$\begin{array}{l}\displaystyle \frac{n{({\rm{C}}{{\rm{H}}}_{3}{\rm{C}}{\rm{O}}{\rm{O}})}_{2}{\rm{Z}}{{\rm{n}}}_{{\rm{KeqE}}}}{n{({\rm{C}}{{\rm{H}}}_{3}{\rm{C}}{\rm{O}}{\rm{O}})}_{2}{\rm{Z}}{{\rm{n}}}_{{\rm{t}}{\rm{o}}{\rm{t}}{\rm{a}}{\rm{l}}}}\approx {\rm{constant}},\\ \,{\rm{when}}\,n{{\rm{H}}}_{2}{\rm{O}}={\rm{constant}},\end{array}$

where:

  • −  
    ${n}{({\rm{C}}{{\rm{H}}}_{{\rm{3}}}{\rm{C}}{\rm{O}}{\rm{O}})}_{{\rm{2}}}{\rm{Z}}{{\rm{n}}}_{{{K}}_{{\rm{e}}{\rm{q}}{\rm{E}}}}$ is the quantity of reacted moles of zinc acetate until the moment of decomposition of the intermediate (achievement of KeqE);
  • −  
    ${n}{({\rm{C}}{{\rm{H}}}_{{\rm{3}}}{\rm{C}}{\rm{O}}{\rm{O}})}_{{\rm{2}}}{\rm{Z}}{{\rm{n}}}_{{\rm{t}}{\rm{o}}{\rm{t}}{\rm{a}}{\rm{l}}}$ is the quantity of zinc acetate in the precursor for the duration of 0 min;
  • −  
    nH2O is the water content in the precursor for the duration of 0 min.

In other words, when increasing the zinc acetate concentration, we should observe only an increase in the quantity of the obtained ZnO NPs of the same size. In order to verify this thesis, we carried out syntheses of ZnO NPs for 5 different concentrations of (CH3COOZn)2 · 2H2O with the constant H2O content in the precursor being 1.5% (table 9). The syntheses of ZnO NPs lasted for 25 min at the temperature of 190 °C. We did not observe significant changes in the average particle size, density, and SSA of ZnO NPs, even when the zinc acetate concentration in the precursor was tripled (table 9). The experimental results confirmed our assumption that an increase in the zinc acetate concentration in the precursor given the 58 times greater quantity of EG would shift the reaction equilibrium towards a greater quantity of products without changes in ZnO NPs size.

Table 9.  Characteristics of synthesis products of NPs ZnO after 25 min of synthesis with different zinc acetate concentration in the precursor.

Sample Cm(CH3COOZn)2 · 2H2O Specific surface area as (m2 g−1) Skeleton density ρs ± σ (g cm−3) Average particle size from SSA BET, d ± σ (nm) Average crystallite size from Nanopowder XRD Processor Demo, d ± σ (nm) Average crystallite size, Scherer's formula, da, dc (nm)
ZnO-1.5%H2O 0.1215 39.9 5.26 ± 0.03 29 ± 1 23 ± 6 22a, 25c
ZnO-1.5%H2O 0.1822 41.1 5.25 ± 0.02 28 ± 1 22 ± 6 21a, 24c
ZnO-1.5%H2O 0.2430 42.3 5.27 ± 0.02 27 ± 1 23 ± 6 22a, 26c
ZnO-1.5%H2O 0.3037 38.5 5.25 ± 0.02 30 ± 2 25 ± 7 24a, 29c
ZnO-1.5%H2O 0.3644 40.6 5.29 ± 0.02 28 ± 3 25 ± 7 24a, 29c

(3) The suggested synthesis mechanism explains the ZnO NPs size control through the shift of the equilibrium quantity of esterification products. The shift of the equilibrium quantity of esterification products depends on two factors: concentration of reagents (mainly H2O, EG) and the value of the equilibrium constant of a given esterification reaction KeqE. If the suggested ZnO synthesis mechanism is correct, then in line with the change of value of the equilibrium constant KeqE, when reagent concentration is maintained constant, different equilibrium quantities of products will be obtained, and as a consequence different ZnO NPs sizes. In order to change the value of esterification equilibrium constant, a different by-product—an ester—should be obtained. This is possible only by changing the solvent (polyol) and at the same time the substrate of the ZnO synthesis reaction.

In order to verify this experimentally, we carried out syntheses of ZnO NPs in EG, DEG, TEG and TTEG, with the constant content of zinc acetate (Cm = 0.1215 mol dm−3) and H2O (1.5%). ZnO NPs syntheses lasted for 25 min at the temperature of 190 °C. The characteristics of the obtained ZnO NPs is included in table 10. The average size of the obtained ZnO NPs in EG was 29 nm, in DEG—32 nm, in TEG—44 nm, while in TTEG—47 nm. Figures 16 and 17 show the change of size and shape of ZnO caused by the use of different polyols. Spherical ZnO NPs were obtained in EG and DEG, while the rod-like shape of ZnO NPs was obtained in TEG and TTEG. The change of shape of the obtained ZnO NPs may result from the solvent's impact on the direction of crystal growth through e.g. the spherical effect.

Table 10.  Characteristics of ZnO NPs samples obtained in different polyols for synthesis duration of 25 min and temperature of 190 °C.

Sample ZnO-EG ZnO-DEG ZnO-TEG ZnO-TTEG
Cm(CH3COOZn)2·2H2O (mol dm−3) 0.1215 0.1215 0.1215 0.1215
H2O content (wt%) 1.50 ± 0.01 1.51 ± 0.01 1.52 ± 0.04 1.51 ± 0.03
Specific surface area, as (m2 g−1) 39.9 36.0 26.4 24.6
Skeleton density, ρs ± σ (g cm−3) 5.26 ± 0.03 5.24 ± 0.03 5.32 ± 0.03 5.38 ± 0.03
Average particle size from SSA BET, d (nm) 29 ± 1 32 ± 1 43 ± 1 47 ± 2
Average crystallite size, Scherer's formula, da, dc (nm) 22a, 25c 27a, 47c 32a, 55c 34a, 63c
Average crystallite size from Nanopowder XRD Processor Demo, d ± σ (nm) 23 ± 6 33 ± 11 39 ± 14 46 ± 17
Figure 16.

Figure 16. Crystallites size distributions of ZnO NPs samples, obtained with the use of Nanopowder XRD Processor Demo.

Standard image High-resolution image

The obtained results confirmed our assumptions regarding the ZnO particle size control through a change of polyol (KeqE) with the constant H2O content. Similar correlations between the change of ZnO NPs size were found by Chieng et al [18], but the average size of ZnO NPs in EG obtained by them was 20 nm, in DEG—39 nm, while in TTEG 69 nm. Chieng et al interpreted that the change of the particle size resulted from the influence of the change of glycol chain length. It should be pointed out that the greater differences in the average size of ZnO NPs obtained by them result from additional changes of H2O content in precursors. Chieng et al did not control changes of H2O quantity in precursors.

(4) In order to confirm that ZnO NPs grow until the substrate (zinc acetate or its derivatives) is exhausted, we carried out ZnO NPs syntheses with the following parameters: temperature >300 °C, pressure 35 bar, duration 41 min. Figure 18(a) presents the obtained ZnO NPs suspended matter after the sedimentation process. The black liquid visible above the white deposit is degraded EG. The change of EG colour proved the occurrence of degradation process caused by too high temperature and too long heating duration. However, despite achieving a temperature exceeding 300 °C with the synthesis duration of 41 min, the average ZnO NPs size was 22 ± 6 nm (figure 18(b), table 11). Figure 19 shows the spherical shape of those ZnO NPs. For the ZnO NPs sample obtained for the synthesis duration of 25 min and temperature of 190 °C, the average particle size was 23 ± 6 nm. The low density of ZnO NPs-41 min being only 4.92 g cm−3 resulted from the presence of impurities coming from EG degradation (table 11). The greater value of SSA in comparison with the sample obtained at the temperature of 190 °C confirms the modification of its surface by the degradation products (soot, tar). In the microwave solvothermal synthesis of ZnO NPs, EG fulfils the role of the solvent, substrate and the stabilising medium, which blocks the growth of ZnO NPs after the exhaustion of substrates. The results of the experiments revealed the lack of increase in ZnO NPs size arising e.g. from the Ostwald ripening process.

Figure 17.

Figure 17. SEM images of ZnO NPs obtained in different solvents (a) EG; (b) DEG; (c) TEG; (d) TTEG.

Standard image High-resolution image
Figure 18.

Figure 18. (a) Image of post-reaction suspended matter of ZnO NPs sample (35 bar, 41 min); (b) crystallite size distribution of ZnO+1.5%H2O, obtained with the use of Nanopowder XRD Processor Demo.

Standard image High-resolution image

Table 11.  Characteristics of the obtained ZnO NPs samples.

Sample Specific surface area, as (m2 g−1) Skeleton density, ρs ± σ (g cm−3) Average particle size from SSA BET, d ± σ (nm) Average crystallite size from Nanopowder XRD Processor Demo, d ± σ (nm) Average crystallite size, Scherer's formula, da, dc (nm)
ZnO-1.5%H2O-25 min, 190 °C 51.9 5.29 ± 0.03 22 ± 1 23 ± 6 21a, 26c
ZnO-1.5%H2O-41 min, T > 300 °C 57.5 4.92 ± 0.05 21 ± 1 22 ± 6 19a, 26c
Figure 19.

Figure 19. SEM images of ZnO-1.5%H2O-41 min, T > 300 °C sample.

Standard image High-resolution image

The obtained results of the experimental verification of the developed mechanism model confirmed its correctness. Based on the presented synthesis mechanism of ZnO NPs obtained as a result of the reaction of zinc acetate with EG, the results of works by various research groups related to ZnO NPs size control can be explained by the change of:

  • −  
    concentration of reagents, (change of H2O content in the precursor [16, 19]);
  • −  
    synthesis duration, (ZnO growth as a result of presence of unreacted substrates in the precursor [15, 63]);
  • −  
    synthesis temperature (change of kinetics of the reaction of formation of intermediates and change of kinetics of ZnO NPs growth [15, 63];
  • −  
    organic solvent, (different values of esterification equilibrium constants and different H2O contents in the used solvents [1618, 63]);
  • −  
    precursor storage conditions, (changes of H2O content in the precursor [19]).

5. Conclusions

A model of size control mechanism of ZnO NPs formed as a result of the reaction of (CH3COO)2Zn with EG in the presence of H2O was presented and verified.

As a result of zinc acetate hydrolysis, water leads to the formation of acetic acid, which participates in the esterification reaction with EG during the microwave solvothermal synthesis. The esterification reaction products are esters and water. The course of the reaction of obtaining and growth of the intermediate, Zn5(OH)8(CH3COO)2 · xH2O, is possible, in turn, only through the co-existence of the esterification reaction. Only water formed in the esterification reaction takes part in the reaction of obtaining/growth of the intermediate, Zn5(OH)8(CH3COO)2 · xH2O. Once the equilibrium constant of the esterification reaction is achieved, the intermediate rapidly decomposes to ZnO NPs, H2O and esters.

The particle size control arising from the change of water content in the precursor is a consequence of the change of quantity of the formed crystalline nuclei of ZnO (NPs) to the remaining unreacted quantity of the substrate (zinc acetate). After the decomposition of the intermediate to homogeneous nuclei of ZnO (NPs), no subsequent crystal nuclei of ZnO (NPs) are formed as a result of further reactions. The only phenomenon that might occur is growth of the existing nuclei of ZnO (NPs) until the moment of exhaustion of substrates that have not been reacted so far.

In the described solvothermal synthesis reaction of ZnO NPs, water acts as the catalyst. It participates in the reaction with the substrates, forming an unstable intermediate, Zn5(OH)8(CH3COO)2 · xH2O, which is at the same time the catalyst of the esterification reaction.

The microwave solvothermal synthesis of ZnO NPs may occur with an efficiency of 94%–98% only because the reactions leading to the formation of the intermediate and to the growth of ZnO NPs are esterification reactions occurring in the presence of catalysts (Zn5(OH)8(CH3COO)2 · xH2O, ZnO).

Acknowledgments

The paper was prepared as a result of execution of 'PRELUDIUM 10' research project, ref. no UMO-2015/19/N/ST5/03668, financed by the National Science Centre, Poland. A part of the research was carried out with the use of equipment funded by 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 like also to thank S Stelmakh, J Mizeracki and A Presz from the Institute of High Pressure Physics of the Polish Academy of Sciences.

Conflict of interests

The authors declare that there is no conflict of interests regarding the publication of this paper.

Please wait… references are loading.
10.1088/1361-6528/aaa0ef