S. Akbar*ab,
S. K. Hasanain†
b,
O. Ivashenko‡
a,
M. V. Dutkaa,
N. Akhtar§
a,
J. Th. M. De Hossona,
N. Z. Alicd and
P. Rudolfa
aZernike Institute for Advanced Materials, University of Groningen, Nijenborgh 4, NL-9747AG Groningen, The Netherlands. E-mail: sadafakbarmp@yahoo.com
bDepartment of Physics, Quaid-i-Azam University, Islamabad, Pakistan
cNational Centre for Physics, Quaid-i-Azam University Campus, 45320 Islamabad, Pakistan
dBAM Federal Institute for Materials Research and Testing, Richard-Willstaetter-Strasse 11, Berlin, Germany
First published on 30th January 2019
We report on the ferromagnetism of Sn1−xZnxO2 (x ≤ 0.1) hierarchical nanostructures with various morphologies synthesized by a solvothermal route. A room temperature ferromagnetic and paramagnetic response was observed for all compositions, with a maximum in ferromagnetism for x = 0.04. The ferromagnetic behaviour was found to correlate with the presence of zinc on substitutional Sn sites and with a low oxygen vacancy concentration in the samples. The morphology of the nanostructures varied with zinc concentration. The strongest ferromagnetic response was observed in nanostructures with well-formed shapes, having nanoneedles on their surfaces. These nanoneedles consist of (110) and (101) planes, which are understood to be important in stabilizing the ferromagnetic defects. At higher zinc concentration the nanostructures become eroded and agglomerated, a phenomenon accompanied with a strong decrease in their ferromagnetic response. The observed trends are explained in the light of recent computational studies that discuss the relative stability of ferromagnetic defects on various surfaces and the role of oxygen vacancies in degrading ferromagnetism via an increase in free electron concentration.
Initial research efforts in this field were focused on magnetic transition metal (TM) doped SnO2 nanoparticles and thin films (Co, Cr, Mn, Fe, Ni and V)14–16 that display ferromagnetism. To avoid magnetic metal clusters or secondary phases of SnO2 doped with nonmagnetic (NM) elements (e.g. Cu and Zn),4–6 alkali metals (Li and K),7–9 alkali earth metals (Mg),5 non-metals (C and N),10,11 and poor metals (In and Ga)12,13 have also been studied and the FM has been reported. Density functional studies17 have shown that Sn vacancies (VSn) are responsible for the observed giant magnetic moment (GMM) of TM-doped SnO2. Other computational studies18 describe surface magnetism induced in a C-doped (001) surface and the incorporation of Li1+ at (001) surface sites19 of SnO2. Surface magnetism in Cu-doped (110) surfaces in SnO2 thin films has also been predicted.20 The role of divalent zinc as a substituent for Sn4+ is particularly interesting due to the closeness in their respective ionic sizes on the one hand and the difference between their respective valences, on the other. Doping of SnO2 with Zn has been shown to induce magnetism in nanoscale systems,6,21 while computational studies performed on the bulk SnO2 system with Zn doping relate this magnetism to the native defect of tin vacancies.22,23 The other prevalent defects in this system, namely oxygen vacancies (VO) are known to weaken FM. It has been reported24 that divalent Zn2+ and Cd2+ ions substituting for Sn4+ introduce holes in the 2p orbitals of the O atoms while the induced magnetic moment arises mainly from the O 2p orbitals and is largest at the first O atom neighbouring the dopant. Hence electron deficiency at the oxygen site, whether originating from the VSn or from replacement of Sn4+ by Zn2+, leads to holes in the O 2p band and to the possible polarization of this band. According to these studies24 the contribution of the VSn defect to the moment itself is usually very small.
The incorporation of zinc in SnO2 and the stability of other common defects is however sensitive to the specific surface where these defects formed. Pushpa et al.25 have studied the formation energies and magnetic moments for various defects in the bulk, at the surface and in sub-surface layers, in both Sn-rich and O-rich conditions. In general it is easier to form both Sn and O vacancies at surfaces than in the bulk, and the (001) surface is preferred to the (110) surface. These authors have shown that although the VSn defect is magnetic for both bulk and (110) surfaces, its formation energy remains very high even at the surface. The oxygen vacancy on the other hand has no magnetic moment in the bulk nor on either of the two surfaces studied. The Zn substitutional defect (ZnSn) possesses a small moment of ∼0–0.11 μB/Zn, but generates a significant moment per unit cell (2 μB per cell). The formation energy for ZnSn on the (001) surfaces is about half that of the bulk, while on the (110) surfaces it is close to that of the bulk. Interestingly, the Zn atom on the (110) surface does not contribute to the induced moment directly.25 The moment arises from the nearest neighbour bridging oxygen and minimally from the in-plane oxygen or the Sn atom. It is further understood that while the moment arises from the polarization of the partially filled oxygen bands, the occurrence of ferromagnetism or antiferromagnetism (AFM) in these bands depends on the separation between the holes on oxygen atoms surrounding the VSn defect.
Alongside these studies of the electronic and FM/AFM properties, there are various reports of unique morphologies in SnO2-based systems that display hierarchical nanostructures with nanorods, nanosheets and nanoflowers at different scales, often as a function of stoichiometry.26–28 The formation of these nanostructures is related to differences in the growth rates of various crystallographic planes in the presence of these defects, e.g. oxygen vacancies, substitutional atoms, etc. Several previous reports26,27,29–34 have demonstrated that the morphologies and properties of SnO2 can be modified by Zn doping since incorporation of zinc into the SnO2 lattice modifies the local structure and the growth rate of different crystallographic planes. In particular Zn2+ ions in the SnO2 lattice inhibit the growth along the [110] direction, promoting the anisotropic growth of nanorods. Because the formation energies of the moment supporting defects (e.g. Sn vacancies) are different for different such surfaces or planes, the morphological and electronic properties become interrelated in this system. While this aspect may not be particularly important for bulk systems, it assumes a different significance for nanostructured materials, where preferential growth of a certain surface affects the magnitude of the magnetic moment.
In the light of the preceding discussion on the role of specific planes in lowering the formation energy of defects and stabilizing ferromagnetism, we prepared nanoparticles of Zn-doped SnO2. A solvothermal synthesis route was adopted that resulted in hierarchal nanoarchitectures with well-defined planar surface structures that change with Zn concentration. Alongside the structural changes the magnetic behaviour also changes and this study contributes to the deeper understanding of the observed variations in structure, morphology and magnetic property relationship.
A PANalytical X'Pert PRO X-ray diffractometer (XRD) equipped with Cu Kα radiation (λ = 1.5405 Å) was used for the structural analysis of the samples. The morphology and microstructure of the samples were investigated by Field Emission Scanning Electron Microscopy (XL30-FEI ESEM-FEG, 5k–30 kV), equipped with energy-dispersive X-ray spectroscopy (EDS) and High Resolution Transmission Electron Microscopy (HRTEM) (JEOL2010FEG operating at 200 kV). X-ray photoelectron spectroscopy (XPS) was employed to analyse the chemical composition of the prepared Zn-doped SnO2 nanoparticles. XPS data were collected using a Surface Science SSX-100 ESCA spectrometer equipped with a monochromatic Al Kα X-ray source (hν = 1486.6 eV) under UHV conditions. All binding energies were referred to the C 1s line at 284.6 ± 0.1 eV (stemming from adventitious carbon). Spectral analysis comprised a Shirley background subtraction and peak deconvolution employing a convolution of Gaussian and Lorentzian functions with a 90–10% ratio by a least-square fitting program (Winspec), developed in the LISE laboratory of the Facultés Universitaires Notre-Dame de la Paix, Namur, Belgium. Magnetic characterization of the samples was carried out using a Quantum Design MPMS-XL7 SQUID magnetometer.
A typical XRD pattern along with the refinement is shown in Fig. 1(b) for the Sn0.96Zn0.04O2 sample. It can be clearly seen that the experimentally observed X-ray peaks accurately match with simulated pattern refined on mineral Cassiterite model in tetragonal space group P42/mnm. Using the XRD data, the cell parameters a and c were calculated for different doping concentrations (x), and their average values are plotted in Fig. 2(a) and (b) respectively. The trend of the calculated values clearly indicates an increase in the values of lattice parameters a and c with increasing Zn concentration up to x = 0.04. The observed expansion could be due to interstitial occupation however we understood this expansion as substitution of Zn2+ with larger ionic radii of 0.74 Å at Sn4+ (0.69 Å) ions. This indicates that the Zn dopant atoms are accommodated substitutionally, filling tin vacancies. In addition, the substitution of Sn-ions by Zn can generate oxygen vacancies for charge compensation.6 Zn ions in solid solution can be excluded because we do not observe maximum full width at half maximum (FWHM) at x = 0.04.
![]() | ||
Fig. 2 Variation of (a) lattice parameter c and (b) lattice parameter a determined from XRD as function of Zn concentration (x) in Sn1−xZnxO2. |
For x > 0.04 both a and c decrease up to the highest concentration studied, namely x = 0.10. This decrease on substitution would also lead to a cell volume reduction since the size of Zn2+ is much smaller than that of O2− (1.4 Å). We also note that with increasing Zn concentration, the diffraction peaks decrease in intensity and tend to become broader as shown in the inset of Fig. 1(a). The changes in intensity and full width at half maximum (FWHM) indicate that the incorporation of Zn dopants results in the deterioration of crystallinity and the decrease of grain size in Sn1−xZnxO2 samples. The average grain size was estimated using the FWHM of (110) and (101) peaks based on Scherrer's equation. As the Zn concentration in Sn1−xZnxO2 increases from x = 0.02 to x = 0.10, the average grain size decreases from 80.0 ± 2.1 nm to 15.0 ± 2.1 nm. This aspect will be discussed further in the context of the morphological studies on these particles.
Energy dispersive X-ray spectroscopy (EDS) allows to check the presence of any unwanted magnetic impurity within the instrumental detection limit of 1%. The analysis confirms that there are no detectable traces of magnetic impurities in the compounds.
The results are shown in Fig. 3. The elemental analysis corroborates the presence of Zn, Sn and O as well as giving evidence for Si and C signals coming from the sample holder with conductive tape.
![]() | ||
Fig. 3 Energy-dispersive X-ray spectra of (a) undoped, (b) Sn0.96Zn0.04O2 and (c) Sn0.90Zn0.10O2 nanoparticles. |
Fig. 4(c) and (d) present images for the Sn1−xZnxO2 with x = 0.06 and x = 0.10 respectively. With increasing Zn concentration the spherical and cubic particles become eroded and acquire different shapes of hierarchical structures. These include a mixture of hemi- and hollow spheres, and elongated chains of flower-type structures. We also observed that for higher zinc concentration these μm-size structures become interconnected by nanoneedles on their surfaces. For x = 0.10 one sees a clear erosion of the individual cubical and/or spherical structures, which now agglomerate to form a bundle of nanoflowers or different shapes, with nanoneedles on their surfaces. For further analysis of the nanoneedles, TEM and HRTEM micrographs were collected from a Sn1−xZnxO2 sample with x = 0.04.
Fig. 5(a) presents an image of this sample where μm-sized particles with mainly cubical shapes can be seen. High resolution TEM images in Fig. 5(b–e) clearly show that these μm-size particles are covered with outward growing nanoneedles, nanorods and nanostructures extending from the surface. The length of these nanoneedles is in the range 10–100 nm (Fig. 5(d) and (e)), and connecting nanorods measure about 85 nm × 280 nm (Fig. 5(d)). Fig. 5(f) and (g) show HRTEM micrographs of long and small nanoneedles. The two groups of crystallographic planes marked in the images have interplanar distances of 0.35 nm and 0.26 nm respectively. These separations match well with the (110) and (101) planes of rutile SnO2.
![]() | ||
Fig. 5 TEM (a–e) and HRTEM (f and g) images of ZnxSn1−xO2 with x = 0.04. HRTEM focuses on the nanoneedles extending from the nanoparticle surfaces in (a–e). The (110) and (101) planes are visible. |
In SnO2 the (110) surface has the lowest surface energy, followed by the (100), (101) and (001) surfaces.40 Nanocrystals have a high surface-to-volume ratio and tend to aggregate to decrease the surface energy. During the initial stages of the solvothermal process, SnO2 spherical nanoparticles were produced with diameters in the range 50–200 nm,41 at higher Zn concentrations these nanoparticles aggregate for energy minimization into solid cubes with needles on the surface. It should be noted that these solid cubes and spheres are all composed of nanocrystalline particles, as shown in Fig. 5(b). A similar behaviour has been demonstrated in the preparation of other hollow structures, such as hollow Cu2O cubes and hollow TiO2 spheres.42,43 Furthermore, in the absence of Zn2+ ions, a similar morphology was not obtained and pure SnO2 (Fig. 4(a)) shows no evidence for nanoneedle-like growth or interconnecting nanorods. While similar nanostructures have been reported in pure SnO2 nanoparticles, their development in un-doped SnO2 appears connected to the presence of Sn2+ ions. In the case of Zn-doped SnO2, the role of Zn2+ as a structure directing agent has been reported26 and is confirmed by our results.
The substitution of Zn2+ for Sn4+ leads to doubly charged oxygen vacancies, as a charge compensation mechanism.6 Consequently the charge density and surface energy of various crystal faces is changed, leading to a large polarity in the growth of Zn-doped SnO2, which in turn yields different growth rates for different surfaces. As already mentioned the average crystallite size estimated by Scherrer's formula decreases with the addition of Zn, suggesting that the zinc dopant plays an active role in reducing the average crystallite size.
After nucleation most of the Zn2+ ions will segregate to surface/interface sites because of the abundant surface area available.34 When these ions occupy the surface sites of SnO2 nanocrystals, they most likely inhibit the formation of necks between particles and the coalescence of tiny SnO2 crystals into larger particles (microcubes and microspheres). Thus our study helps establish some important differences between the morphologies of low and high zinc concentration nanoparticle systems. Whereas ZnxSn1−xO2 particles with x = 0.04 show well-developed isolated structures with needle-like growth on the surfaces, particles with higher Zn concentration (x = 0.06, 0.10) exhibit small grain sizes, erosion of shapes and agglomeration of the particles.
The undoped SnO2 spectrum shows Sn 3d5/2 and 3d3/2 core levels of the Sn4+ ions at binding energies of 486.9 and 495.3 eV respectively. For x = 0.04 the binding energy of Sn 3d doublet (485.9 eV and 494.3 eV) decreases by 1 eV and for x = 0.10 by 1.7 eV (485.2 and 493.7 eV) as compare to undoped SnO2. This decreases in binding energy of the Sn 3d doublet can be attributed to the presence of oxygen vacancies with addition of Zn dopant.31,44,45 It can be noted that at x = 0.10 the chemical shift in Sn 3d is not following the same trend as it is showing from x = 0.00 to x = 0.04 which could be due to the reason that at higher concentration Zn going more to bulk sites as compare to surface sites.
Fig. 6(b) displays the O 1s spectra for undoped, x = 0.04 and x = 0.10 Zn doped SnO2 samples. The main peak (centred at 531.2 eV for x = 0.00, at 529.9 eV for x = 0.04 and at 528.6 eV for x = 0.10) was assigned to the coordination of oxygen in Sn–O–Sn, while the shoulder at higher binding energy side could be ascribed to Sn–O–Zn bonds.31,39 The chemical shift towards lower binding energy as a function of Zn-doping can be attributed to the increasing number of VO. To analyse it further, Fig. 6(d) presents the fit of the O 1s spectra of x = 0.04 and 0.10. The deconvolution requires two components namely OI and OII. OI centred at 529.9 eV and 528.4 eV, and OII, centred at 531.7 eV and 529.9 eV correspond to lower and higher binding energy components for x = 0.04 and 0.10 respectively. The lower binding energy component OI is attributed to the coordination of oxygen bound to Sn atoms, whereas the higher binding energy component OII is assigned to the oxygen vacancies. The OII component is larger for x = 0.10 (26%) than for x = 0.04 (16%), indicating that the number of oxygen vacancies increases with the zinc concentration.
Fig. 6(c) displays the Zn 2p core level region where the Zn 2p3/2 peak appears at a binding energy of 1021.1 eV,21,46 confirming that Zn atoms were incorporated into the Sn lattice and form Zn–O bonds. We analysed at least three different spots on every sample and found that the Zn concentration did not vary between different spots, which indicates that the Zn concentration in these samples is homogeneously distributed. We find out Zn/Sn ratio by XPS data. For x = 0.04 the ratio came out as expected, however a lower number was obtained for x = 0.10 i.e. 0.05. This could be understood that Zn goes to surface first for lower concentration, however with the increasing concentration of dopants, Zn also starts going into the bulk and XPS shows lower number as being more surface sensitive technique.
It can be concluded that at low concentrations Zn substitutes more to surface sites than at bulk. It is well known in nanoparticles, that surface sites are more activated.34,49
The x = 0.04 (inset Fig. 7(a)) and 0.06 compositions were also measured at low temperature and showed a strong increase in both the remanence and hysteresis. The second component of the magnetization, as is evident from Fig. 7(a), is a linearly increasing or paramagnetic component.
The magnetization data for x = 0.04 as a function of temperature for an applied field of 1000 Oe is shown in Fig. 7(b), while the inset shows the inverse of magnetization versus temperature. It is apparent from the curvature of the 1/M versus T that the data does not show a good Curie–Weiss behaviour. This is of course not surprising, since the full magnetization includes a ferromagnetic component in addition to the paramagnetic part.
To separate the two components, we fitted the higher field data of the magnetization versus magnetic field (Fig. 7(a)) to the expression M = Mo + eH (where e is a constant) and from the linear fit the value of Mo was extracted, which will be referred to as the ferromagnetic component. The procedure is illustrated in Fig. 8(a). The values of e, the fitting constant representing the paramagnetic susceptibility, were obtained from the fits at T = 50 K and 300 K respectively. The ratio was found to be 5.2, which is close to the expected value of 6 for a purely paramagnetic behaviour, χ ∼ 1/T. The ferromagnetic component, Mo, for each composition was obtained by the above procedure and then subtracted from the full moment measured at H = 10 kOe. The resultant value for each composition is referred to as paramagnetic component. The same procedure was also followed for the analysis of the data at lower temperatures, for Sn1−xZnxO2 with x = 0.04 and x = 0.06.
The variation of both the magnetic components is shown as a function of the composition in the main part of Fig. 8(b). The inset shows the same two components with the magnetization in Bohr magnetons per zinc atom. Here the number of zinc atoms corresponds to the nominal concentration. We can see that the ferromagnetic moment per zinc atom has a maximum at x = 0.04 and is somewhat smaller for the x = 0.02, while it falls sharply at x = 0.06. We note that for x = 0.02 and 0.04 the lattice constant shows an expansion compared to x = 0.00, which indicates that zinc substitutes at Sn sites. The strong decline of the ferromagnetic moment/zinc atom for x = 0.06 coincides with the contraction of the lattice constant as shown in the XRD data.
Similarly the paramagnetic component/zinc atom (inset of Fig. 8(b)) is maximum at x = 0.04 and decreases very little for x = 0.06. However this contribution drops very strongly for Sn1−xZnxO2 with both x = 0.02 and x = 0.10. This suggests that while the zinc dopants do not lead to a large ferromagnetic component for x > 0.04, they are still able to contribute strongly to the paramagnetic part for x = 0.06.
Consistent with the earlier discussion of the structural and electronic effects (see discussion of XPS data) it is possible that the observed variation of the ferromagnetic moment per zinc atom reflects a competing effect of two contributions, namely zinc dopants as hole contributors and stabilizers of Sn vacancies on the one hand, and oxygen vacancies as electron donors on the other.
For x = 0.02 and 0.04 the zinc atoms substitute for Sn but the number of oxygen vacancies is relatively small, leading to a larger ferromagnetic component. For higher Zn content it is possible that in addition to increasing the number of O vacancies, zinc occupies some sites other than those of Sn, e.g. O sites. Both features are expected to lead to a decline of the ferromagnetic behaviour. However the paramagnetic part can still be contributed to, as will be discussed next. There may be two different sources for this paramagnetic part. Firstly, the magnetic moment developed at sufficiently isolated defect sites (ZnSn) will not lead to a stabilization of the FM17 but the paramagnetic contribution may still exist. Secondly we note that singly ionized oxygen vacancies can also yield a paramagnetic contribution. At room temperature all neutral VO centres (an oxygen vacancy with two trapped electrons) are dissociated into a singly ionized oxygen vacancy (an oxygen vacancy with a single trapped electron) and a free electron. These singly charged oxygen vacancies have been reported to be paramagnetic.47 The decrease in this paramagnetic contribution at higher zinc concentration (x = 0.10) is then attributable to the recombination of O 2p holes with the trapped electron of the
, converting it47,48 into a nonmagnetic
We believe that the latter is the most plausible explanation for the variation of the paramagnetic moment with zinc concentration.
Footnotes |
† Now at COMSTECH Secretariat, 33-Constitution Avenue, G-5/2, 44000 Islamabad, Pakistan. |
‡ Now at Centre for Materials Science and Nanotechnology, University of Oslo, Sem Sælands vei 26, Kjemibygningen, 0371 Oslo, Norway. |
§ Now at Department of Physics and Technology, University of Bergen, Bergen, Norway. |
This journal is © The Royal Society of Chemistry 2019 |