Manipulating coupling state and magnetism of Mn-doped ZnO nanocrystals by changing the coordination environment of Mn via hydrogen annealing
Cheng Yan 1 , Li Wen-Xian 2, 3 , Hao Wei-Chang 1, †, , Xu Huai-Zhe 1, ‡, , Xu Zhong-Fei 1 , Zheng Li-Rong 4 , Zhang Jing 4 , Dou Shi-Xue 3 , Wang Tian-Min 1
Department of Physics and Center of Materials Physics and Chemistry, Beihang University, Beijing 100191, China
Solar Energy Technologies, School of Computing, Engineering and Mathematics, University of Western Sydney, Penrith, NSW 2751, Australia
Institute for Superconducting and Electronic Materials (ISEM), University of Wollongong, Wollongong, NSW 2522, Australia
Beijing Synchrotron Radiation Facility, Institute of High Energy Physics, Chinese Academy of Sciences, Beijing 100049, China

 

† Corresponding author. E-mail: whao@buaa.edu.cn

‡ Corresponding author. E-mail: hzxu@buaa.edu.cn

Project supported by the National Basic Research Program of China (Grant No. 2013CB934001) and the National Natural Science Foundation of China (Grant Nos. 51072012 and 51272015).

Abstract
Abstract

Mn-doped ZnO nanocrystals are synthesized by a wet chemical route and treated in H 2 /Ar atmosphere with different H 2 /Ar ratios. It is found that hydrogen annealing could change the coordination environment of Mn in ZnO lattice and manipulate the magnetic properties of Mn-doped ZnO. Mn ions initially enter into interstitial sites and a Mn 3+ O 6 octahedral coordination is produced in the prepared Mn-doped ZnO sample, in which the nearest neighbor Mn 3+ and O 2 ions could form a Mn 3+ –O 2− –Mn 3+ complex. After H 2 annealing, interstitial Mn ions can substitute for Zn to generate the Mn 2+ O 4 tetrahedral coordination in the nanocrystals, in which neighboring Mn 2+ ions and H atoms could form a Mn 2+ –O 2− –Mn 2+ complex and Mn–H–Mn bridge structure. The magnetic measurement of the as-prepared sample shows room temperature paramagnetic behavior due to the Mn 3+ –O 2− –Mn 3+ complex, while the annealed samples exhibit their ferromagnetism, which originates from the Mn–H–Mn bridge structure and the Mn–Mn exchange interaction in the Mn 2+ –O 2− –Mn 2+ complex.

1. Introduction

Nowadays, semiconductor devices are a great challenge to data processing and storage capacity. Generally, this could be improved by simultaneously manipulating the spin and charge properties of electrons. Diluted magnetic semiconductor (DMS), a promising candidate for spin-polarized-current-injection material, [ 1 , 2 ] especially ZnO has recently attracted extensive attention. [ 3 , 4 ] Massive theoretical studies have been conducted and predicted that p-type Mn-doped ZnO could present ferromagnetic behavior at room temperature (RT). [ 5 7 ] Different magnetic properties in Mn-doped ZnO such as ferromagnetic, [ 8 , 9 ] paramagnetic, [ 10 , 11 ] or antiferromagnetic behavior [ 12 ] have been observed. Zheng et al. [ 12 ] showed that Mn-doped ZnO tetrapod structures have a Curie temperature T C = 43 K, which originates from (Zn, Mn) Mn 2 O 4 at the surface. Cheng and Chien [ 13 ] observed paramagnetic behavior caused by homogeneous distribution of Mn in epitaxial Mn-doped ZnO thin film. Inamdar et al. [ 14 ] found that incorporation of Mn 2+ ions at substitutional sites in perfect crystal produces a higher saturation magnetic moment than in the un-doped material. Through the above researches, we can draw a conclusion that the coordination environment of Mn in ZnO plays a key role in understanding the magnetic mechanism, and there is a fundamental relation between the observed ferromagnetism and substitutional Mn in the ZnO system. It is very important to investigate the coordination environment of Mn in order to clarify the origin of magnetism in the Mn-doped ZnO system.

The oxidation states and the local coordination of the selected atom, even in a diluted context, can be exactly studied by x-ray absorption fine structure spectroscopy (XAFS). [ 15 ] By analyzing the Mn and Zn K -edge XAFS spectra, Céspedes et al. [ 16 ] found that the ferromagnetism in Mn-doped ZnO originated from high-density regions of Mn cations with a MnO 2 – δ distorted environment at non-stoichiometric and highly defective ZnO grain boundaries. Zhang et al. [ 17 ] suggested that the Mn-doped ZnO system exhibited substantial paramagnetic behavior and the strong antiferromagnetism (AFM) demonstrated by the mere occupation of Zn sites with Mn ions, forming spin-antiparallel pairs and inducing a large structural distortion, which was confirmed by the XAFS spectra.

As is well known, hydrogen is inevitably introduced into the ZnO lattice and will form shallow donors. We have obtained that (Mn, N) co-doping can enhance the solubility of Mn in the ZnO lattice via reducing the donor binding energy of Mn impurity by orbital hybridization between the N-acceptor and Mn-donor. [ 18 ] We considered that the donor binding energy of the impurity should decrease if hybridization between H-donors and Mn-donors exists. In addition, some theoretical calculations and experimental results indicated that interstitial hydrogen can induce or enhance ferromagnetism in transition metal (TM) ion doped ZnO. [ 19 21 ] In a hydrogen annealed Mndoped ZnO system, the magnetic properties are different, and the source of the observed magnetism is still under intense debate. Manivannan et al. [ 22 ] noted that the magnetism of Mndoped ZnO is unchanged after hydrogenation. Sharma and Varma [ 23 ] proposed that the RT ferromagnetism in their hydrogenated sample is generated from oxygen vacancies and Zn interstitials. Our previous work [ 20 ] has stated that there is a great enhancement of ferromagnetism in hydrogen annealed Co-doped ZnO nanocrystal due to the formation of stable structure Co–H–Co. We supposed that similar structures could be observed by annealing Mn-doped ZnO in hydrogen, which may cause ferromagnetic coupling in the nanocrystal.

The aim of this work is to establish the relationship between the magnetic properties and the coordination environment of Mn in the ZnO lattice. In this work, we synthesize Mn-doped ZnO nanocrystals by the wet chemical route and treat them in H 2 /Ar atmosphere. The Mn coordination environment is investigated by XAFS. We demonstrate that hydrogen annealing can manipulate the coordination environment of Mn, thus inducing magnetic property variation in Mn-doped ZnO.

2. Experimental section

Mn-doped ZnO nanocrystals were synthesized by the wet chemical route as described previously. [ 18 ] The precursor salts (zinc acetate and manganese acetate) were dissolved in dimethyl sulfoxide (DMSO) according to the desired stoichiometry. The total metal ion concentration was 0.1 M. Tetramethylammonium hydroxide (TMAH) was dissolved in ethanol, keeping the concentration at 0.55 M. Then, the ethanol solution of TMAH was dropwise added into the metal salt solution with constant stirring at room temperature. After 48 h, the nanocrystals were precipitated by adding ethyl acetate and removed from the supernatant by centrifugation. Finally, the precipitates were annealed at 873 K in air for 3 h in order to remove the precursor. The obtained powders were annealed under H 2 /Ar atmospheres with different ratios of H 2 :Ar at 873 K for 2 h. The ratios of H 2 /Ar were 1:5 and 4:5, respectively. We denote the as-prepared and H 2 annealed samples as ZnMnO, ZnMnO-H1 (1:5), and ZnMnO-H2 (4:5), respectively.

X-ray diffraction (XRD) patterns were collected on a Rigaku Dmax/C x-ray diffractometer using Cu K α radiation operated at 40 kV and 200 mA. Mn K -edge x-ray absorption near edge structure (XANES) and extended x-ray absorption fine structure (EXAFS) spectra were collected in fluorescence mode due to the lower concentration of Mn on Beamline 1W1B at the Beijing Synchrotron Radiation Facility and Beamline BL14W on the Shanghai Synchrotron Radiation Facility. Raman spectra were collected on an HR800 LabRam Infinity spectrometer excited by a continuous Ar + laser with a wavelength of 514 nm. The magnetic measurements were carried out using a physical properties measurement system (PPMS; Quantum Design) equipped with a vibrating sample magnetometer.

3. Result and discussion
3.1. Characterization of the coordination environment

Figure 1 shows x-ray diffraction (XRD) patterns of ZnMnO, ZnMnO-H1, and ZnMnO-H2. All the diffraction peaks could be attributed to pure ZnO with a wurtzite structure. No secondary phase is detectable within the resolution of XRD. Figure 2 shows the TEM images of ZnMnO and ZnMnO-H2. There is no obvious variation between these two samples, which indicates that annealing under H2/Ar atmosphere has no effect on the morphology of ZnMnO.

Fig. 1. XRD patterns of ZnMnO, ZnMnO-H1, and ZnMnO-H2.

Figure 3 shows the normalized Mn K -edge XANES spectra of ZnMnO, ZnMnO-H1, ZnMnO-H2, MnO, Mn 2 O 3 , and ZnO reference samples. For comparison between Zn and Mn XANES spectra, the energy is rescaled to the respective absorption K -edges calculated from the first derivative of the XANES signal. [ 24 ] The Mn K -edge energy is used to characterize the average Mn oxidation state. The main edge energy shifts toward lower energy from ZnMnO to ZnMnO-H2, indicating a decrease in the oxidation state with increasing the value of H 2 /Ar. The peaks with positions marked by the dotted lines for our samples are the characteristic peaks of MnO and Mn 2 O 3 , which reveal that Mn oxidation states range between Mn 2+ and Mn 3+ . The valence of Mn in the un-hydrogenated sample is about +3, and ZnMnO-H2 mainly presents Mn 2+ oxidation states, while in ZnMnO-H1, Mn ions are found to be in a mixed valence state of Mn 2+ and Mn 3+ . Furthermore, the difference between the Mn K -edge features of ZnMnO and the Zn K -edge features of ZnO at 10–70 eV indicates that the chemical environment of Mn in ZnMnO is different from that of Zn in ZnO. It implies that the Mn 3+ ions do not substitute for Zn in the ZnMnO sample. When annealing in H 2 /Ar, the similar features at about 10–70 eV between the Mn K -edge of ZnMnO-H1 and the Zn K -edge of ZnO reveal that the chemical environment of some Mn is the same as that of Zn in ZnO. This means that part of the Mn 3+ ions have substituted for Zn 2+ in ZnO. With increasing the value of H 2 /Ar, the features in the post-edge region between the Mn K -edge of ZnMnO-H2 and the Zn K -edge of ZnO are almost the same, which reveals that most of the Mn ions have substituted for Zn.

Fig. 2. TEM images of ZnMnO (a) and ZnMnO-H2 (b).
Fig. 3. Normalized Mn K -edge XANES spectra of ZnMnO, ZnMnO-H1, ZnMnO-H2, and the reference samples of MnO, Mn 2 O 3 , and ZnO.

Figure 4 shows the Mn K -edge Fourier transforms of the k 3 χ ( k ) functions of the EXAFS spectra. For comparison, the Mn K -edges of MnO, Mn 2 O 3 , and MnOOH, as well as Zn K -edge of pure ZnO, are presented along with those of the ZnMnO samples. Here, P 1 (1.41 Å) and P 3 (2.41 Å) are attributed to the first Mn–O and the second Mn–Mn coordination shells of MnO 6 in MnOOH, [ 25 ] respectively. P 2 (1.55 Å) and P 4 (2.93 Å) respectively represent the first Zn–O and the second Zn–Zn shells in ZnO. The coordination of Mn in Mn-doped ZnO turn into a four Mn–O bonded Mn 2+ O 4 tetrahedron while Mn substitutes for Zn. In ZnMnO nanocrystals, there are two peaks located at 1.41 Å and 2.41 Å, which means that the coordination environment of Mn mainly performs as the MnO 6 octahedral configuration. In the ZnMnO-H2 sample, two peaks at 1.55 Å and 2.93 Å indicate that Mn ions are demanded by MnO 4 tetrahedral configuration. In the ZnMnO-H1 sample, three main coordination peaks are found as shown in Fig. 4 . Peak A between peak P 1 and P 2 corresponds to the Mn–O coordination. The other two coexisting peaks, B and C appearing at the positions of P 3 and P 4 , represent the Mn–Mn coordination in MnO 6 and the Mn–Mn(Zn) coordination in MnO 4 , respectively. These illustrate that there are two different configurations of Mn in ZnMnO-H1 nanocrystal. As shown in the inset of Fig. 4 , the calculated areas of peaks B and C according to the Gauss fitting are explained as being due to the fact that the content ratio between MnO 4 and MnO 6 in ZnMnO-H1 is about 4:1.

Fig. 4. Mn K -edge Fourier transforms of k 3 χ ( k ) functions of EXAFS spectra without phase correction for ZnMnO, ZnMnO-H1, and ZnMnO-H2, together with reference samples MnO, Mn 2 O 3 , MnOOH, and ZnO. The inset shows the Gauss fitting of peaks B and C.

Figure 5 shows the room-temperature Raman spectra of our samples. Compared with the spectrum of pure ZnO that we synthesized, we observe six modes at about 99.4, 332.3, 384.4, 438.6, 484.5, and 579.9 cm −1 corresponding to , 2- E 2 (M), , , an interface or surface phonon mode (SPM), and mode, respectively. [ 26 ] Two additional modes at about 526.9 cm −1 and 665.1 cm −1 are denoted as AM1 and AM2. AM1 is assigned to the mode due to Mn 2+ dissolved in the ZnO lattice. [ 27 , 28 ] The origin of AM2 is still under intense debate. Samanta et al. [ 29 ] and Cong et al. [ 30 ] attributed it to the second phases ZnMn 2 O 4 and ZnMnO 3 , respectively. Yet there are no second phases in our samples. In the literature, this mode is reported in the Raman spectra of Mn 2 O 3 [ 31 ] and MnOOH. [ 32 ] As with our samples, the coordination environments of Mn in ZnMn 2 O 4 , ZnMnO 3 , Mn 2 O 3 , and MnOOH are also in the form of a MnO 6 octahedron. As a result, we attribute AM2 to a mode arising from MnO 6 octahedral coordination. As shown in Fig. 5 , the decreasing of the intensity of AM2 results from the decreasing of the concentration of MnO 6 octahedron with increasing the value of H 2 :Ar, which is consistent with the EXAFS spectra.

Fig. 5. Raman spectra of ZnO, ZnMnO, ZnMnO-H1, and ZnMnO-H2.

There are two main positions of Mn in the ZnO lattice after doping. One is interstitial, with the Mn ions forming the MnO 6 octahedral configuration as shown in Fig. 6(a) ; the other is Mn substitution for Zn, forming the MnO 4 tetrahedral configuration (Fig. 6(b) ). The different Mn oxidation states and local environments of our samples are shown in Table 1 . As mentioned above, the coordination of Mn in the as-prepared ZnMnO is primarily that in MnO 6 , which means that most Mn ions are in interstitial sites in the ZnO lattice. In ZnMnO-H2, the coordination of most Mn ions are those in MnO 4 , which is similar to the coordinates of ZnO 4 in ZnO. This reveals that most Mn ions substitute for Zn ions in the ZnO lattice. From the above analysis, we can conclude that hydrogen annealing can control the coordination environment of Mn in the ZnO lattice, i.e., it changes from Mn 3+ O 6 octahedral coordination to Mn 2+ O 4 tetrahedral coordination. As we discussed octylamine treatment of Mn-doped ZnO, [ 18 ] the estimated donor binding energy of an impurity ( E B ) can be expressed as

Here, the dielectric constant, ε , and the charge of the electron, e , are all constant. So E B only depends on the value of the Bohr radius of the donor, a d . In the as-prepared sample, the donor binding energy of the impurity is induced by Mn 3+ , in which there are four electrons in the d-shell. When the sample is annealed in hydrogen, H can offer an electron to the Mn 3+ ion and form a covalent bond with it. The hybridization between Mn and H would enhance the value of a d . According to Eq. ( 1 ), E B would decrease with increasing the value of H 2 :Ar, which induces a transformation of the Mn 3+ O 6 octahedron into Mn 2+ O 4 tetrahedron.

Fig. 6. Schematic diagrams of the interstitial site (a) and the substitutional site (b) of Mn in the ZnO lattice after doping.
3.2. Magnetic properties

Figures 7(a) 7(c) demonstrate the curves of RT magnetization versus field ( M H ). The as-prepared nanocrystal exhibits RT paramagnetism (PM), while the sample annealed in H 2 /Ar atmosphere is ferromagnetic (FM). The values of saturation magnetization ( M S ) of ZnMnO-H1 and ZnMnO-H2 are 0.008 and 0.01 emu/g, respectively. Our previous work [ 33 ] shows that the as-prepared Mn-doped ZnO nanocrystal is antiferromagnetic, and the Neel temperature is below RT. As a result, ZnMnO shows PM at RT. In addition, there is a similar variation between M S and the content of Mn 2+ O 4 tetrahedron, together with the H 2 content as shown in Table 1 . This results in the fact that RT ferromagnetism is related to the Mn 2+ O 4 tetrahedral coordination and H in a ZnO lattice.

Table 1.

Valence, main state of Mn in the ZnO lattice, the local coordination of Mn, and RT magnetic property in ZnMnO, ZnMnO-H1, and ZnMnO-H2.

.
Fig. 7. RT M H curves of ZnMnO (a), ZnMnO-H1 (b), ZnMnO-H2 (c).

We also measure M H curves of ZnMnO-H1 and ZnMnO-H2 at 5 K after cooling from 350 K in 0 and 1000 Oe fields. The loops of zero-field cooling (ZFC) and field cooling (FC) hysteresis are shown in Figs. 8(a) and 8(b) . All the samples are ferromagnetic, and the values of M S are 0.472, and 2.104 emu/g, respectively. In the ZnMnO-H1 sample as plotted in the enlargement near the low-field region in Fig. 8(a) , the hysteresis loop maintains good central symmetry in the ZFC process but displays negative exchange bias behavior in the FC, with the exchange bias field being about 70 Oe in the FC process. As is well known, the exchange bias phenomenon is generally ascribed to the exchange coupling at the interface between the ferromagnetic layer and the antiferromagnetic layer. [ 34 ] It is also reported that there is exchange bias in a single-phase magnet, such as Ru 0.25 Cr 0.75 O, [ 35 ] YbFe 2 O 4 . [ 36 ] As a result, the exchange bias behavior observed at 5 K proves that the system must contain at least two exchange-coupled states, such as antiferromagnetic phase and ferromagnetic coupling. The structural results testify that both Mn 3+ O 6 octahedron and Mn 2+ O 4 tetrahedron exist in the ZnMnO-H1 sample. There is no exchange bias phenomenon in ZnMnO-H2 (Fig. 8(b) ), in which Mn ions mainly exist in the form of Mn 2+ O 4 tetrahedron. Combining the above results and analysis, we believe that antiferromagnetic coupling is induced by Mn 3+ O 6 octahedral coordination, while ferromagnetic coupling is due to Mn 2+ O 4 tetrahedral coordination and some H-related structures. [ 23 ]

Fig. 8. ZFC and FC (for 1000 Oe) M H curves measured at 5 K for (a) ZnMnO-H1 and (b) ZnMnO-H2. The insets show the enlarged low-field regions of these two samples.

Figures 9(a) 9(c) show the variations of magnetization with temperature ( M T curve) of ZnMnO, measured in a temperature range from 5 K to 350 K in a magnetic field of 1000 Oe. In order to further investigate the phase transition, we calculate the inverse magnetic susceptibility as a function of temperature (1/ χ T curve) from M T data. There are magnetic phase transitions at about 100 K in all samples. Martínez et al. [ 37 ] considered it as an AFM–PM phase transition. Linear fitting of the 1/ χ T curve from 5 K to 250 K is carried out in these three samples, and the fitting formulas are χ = 0.71/( T + 39.9), χ = 0.86/( T + 124.5), and χ = 0.9/( T + 34.2), respectively. According to the Curie–Weiss formula, χ = C 0 /( T + θ 0 ), we can obtain the value of the Curie–Weiss temperature, θ 0 , and the Curie constant, C 0 , as shown in Table 2 . The negative value of θ 0 verifies the existence of antiferromagnetic coupling in each of our three samples. The antiferromagnetic behavior could be assigned to the magnetic interaction between trivalent Mn ions in Mn 3+ O 6 octahedral coordination. The Curie constant can be given as

where g is the Lande factor, S is the spin quantum number, μ B is the Bohr magneton, k B is the Boltzman constant, and n is the number of cation sites per unit volume. The exchange integral between the nearest neighbors J can be explained as

where z is the number of nearest-neighbor cations in the case of the wurtzite structure, and θ 0 is the Curie–Weiss temperature. According to Eqs. ( 2 ) and ( 3 ), J can be obtained from

where ( B ) 2 /2 z is constant. So the value of J only depends on the value of 0 / C 0 , with θ 0 and C 0 being evaluated from the extrapolation line, and n being in direct proportion to the number of Mn 3+ . Table 2 summarizes the values of θ 0 , C 0 , n , and J . As the density of Mn 3+ O 6 octahedron decreases, the magnitude of J decreases. This confirms the strong antiferromagnetic exchange coupling in the as-prepared sample and the weak antiferromagnetic coupling in the hydrogen annealed samples. This calculated result provides further evidence to support our hypothesis that antiferromagnetic coupling originates from the Mn 3+ O 6 octahedron.

Fig. 9. (a)–(c) The M T and 1/ χ T curves of ZnMnO, ZnMnO-H1, and ZnMnO-H2 in a temperature ranging from 5 to 350 K in a magnetic field of 1000 Oe.

There is other magnetic phase transition at about 250 K in both hydrogen annealed samples as displayed in Figs. 9(b) and 9(c) . We consider this as an FM–PM transition because the 1/ χ T curve is similar to the scenario in the schematic diagram of the FM–PM phase transition. Annealing in H 2 is an effective method to enhance M S in ZnCoO [ 20 ] or ZnMnO [ 23 ] nanocrystal. Also, this FM–PM phase transition only appears in the annealed sample but not the as-prepared sample, which makes us propose that the ferromagnetic behavior below 250 K could be attributed to some H-related configurations in the hydrogen annealed samples.

Table 2.

Values of θ 0 , C 0 , θ 0 / C 0 , n , and J .

.

In order to establish the relationship between the magnetic coupling and the coordination environment of Mn in the ZnO lattice, crystal lattice and energy level diagrams of a Mn 3+ O 6 octahedron at an interstitial site and a Mn 2+ O 4 tetrahedron at a substitutional site are given in Fig. 10 . In the as-prepared sample (Fig. 10(a) ), the neighboring Mn ions in the Mn 3+ O 6 octahedron form a Mn 3+ –O 2− –Mn 3+ complex with a bond angle of 180°. According to the superexchange interaction, [ 38 ] one p electron of the O atom would couple with the e g orbital of Mn 3+ . In addition, this electron in the p orbital of the O atom must be antiparallel to the e g orbital, which makes the two neighboring Mn 3+ ions antiparallel. Thus, the system exhibits antiferromagnetic behavior. It also explains why most researchers have observed antiferromagnetic behavior in Mn-dopant ZnO. When the nanocrystal is annealed under H 2 , many interstitial H (H i ) atoms are incorporated into the nanocrystal. We hold the opinion that the substitutional Mn (Mn 2+ O 4 tetrahedron) and H i would form a Mn–H–Mn bridge structure. A similar structure has been reported by Park et al. [ 39 ] through electron paramagnetic resonance and nuclear magnetic resonance spectra. In this case, the spins of the two neighboring substitutional Mn ions become parallel via the H i mediator, as shown in Fig. 10(b) . This Mn–H–Mn bridge structure leads to the enhancement of ferromagnetic coupling in hydrogen annealed nanocrystals, as we have observed in hydrogen annealed Co-doped ZnO nanocrystals. [ 20 ] Another reason which gives rise to FM may originate from ferromagnetic coupling to the exchange interaction between two neighboring Mn 2+ ions. In Fig. 10(c) , the bond angle of Mn 2+ –O 2− –Mn 2+ is 120° and the distance between Mn ions decreases when Mn ions substitute for Zn ions. Consequently, the neighboring Mn 2+ ions can couple with each other and a parallel ferromagnetic coupling is produced between the two neighboring Mn 2+ ions. Our results clearly point out the relationship between the magnetic properties and the coordination environment of Mn in the ZnO lattice. This relationship may provide a new perspective for understanding the current controversial experimental results in transition metal doped DMSs. It should be seriously taken into account in future DMS studies.

Fig. 10. Crystal lattice and energy level diagrams of (a) Mn 3+ O 6 octahedron at interstitial site and Mn 3+ –O 2− –Mn 3+ complex, (b) Mn 2+ O 4 tetrahedron at substitutional site and Mn 2+ –H–Mn 2+ complex, (c) Mn 2+ O 4 tetrahedron at substitutional site and Mn 2+ –Mn 2+ complex.
4. Conclusions

We synthesize Mn-doped ZnO nanocrystals by a wet chemical route and post hydrogen annealing. Our results indicate that Mn ions initially enter into interstitial sites in the ZnO lattice and form Mn 3+ O 6 groups with an octahedral coordination in the as-prepared sample. Hydrogen annealing can change the state of Mn into Mn 2+ O 4 , which has a tetrahedral coordination at the substitutional sites. Antiferromagnetism of as-prepared nanocrystals, originating from the interactions of the Mn 3+ –O 2− –Mn 3+ complex in the octahedral coordination, is obtained. The annealed sample exhibits ferromagnetic behavior due to the Mn–H–Mn bridge structure and Mn–Mn complex in the tetrahedral coordination. Our results suggest that the coordination environment of Mn can be changed with hydrogen annealing, which generates a significant effect on the magnetic properties of Mn-doped ZnO. The relationship between the magnetic coupling and Mn coordination environment will provide a new perspective for manipulating magnetic coupling mechanism in DMSs.

Reference
1 Pearton S J Abernathy C R Norton D P Hebard A F Park Y D Boatner L A Budai J D 2003 Mater. Sci. Eng. R 40 137
2 Akinaga H Ohno H 2002 IEEE Trans. Nanotechnol. 1 19
3 Ohno Y Young D K Beschoten B Matsukura F Ohno H Awschalom D D 1999 Nature 402 790
4 Ohno H Chiba D Matsukura F Omiya T Abe E Dietl T Ohno Y Ohtani K 2000 Nature 408 944
5 Dietl T Ohno H Matsukura F Cibert J Ferrand D 2000 Science 287 1019
6 Sato K Katayama-Yoshida H 2001 Physica B 308 904
7 Mukherjee D Dhakal T Srikanth H Mukherjee P Witanachchi S 2010 Phys. Rev. B 81 205202
8 Jung S W An S J Yi G C Jung C U Lee S I Cho S 2002 Appl. Phys. Lett. 80 4561
9 Droubay T C Keavney D J Kaspar T C Heald S M Wang C M Johnson C A Whitaker K M Gamelin D R Chambers S A 2009 Phys. Rev. B 79 155203
10 Fukumura T Jin Z Kawasaki M Shono T Hasegawa T Koshihara S Koinuma H 2001 Appl. Phys. Lett. 78 958
11 Kundaliya D C Ogale S B Lofland S E Dhar S Metting C J Shinde S R Ma Z Varughese B Ramanujachary K V Salamanca-Riba L Venkatesan T 2004 Nat. Mater. 3 709
12 Zheng R K Liu H Zhang X X Roy V A L Djurisic A B 2004 Appl. Phys. Lett. 85 2589
13 Cheng X M Chien C L 2003 J. Appl. Phys. 93 7876
14 Inamdar D Y Lad A D Pathak A K Dubenko I Ali N Mahamuni S 2010 J. Phys. Chem. C 114 1451
15 Steven D C 2000 Los Alamos Sci. 26 422
16 Céspedes E Laguna-Marco M A Jiménez-Villacorta F Chaboy J Boada R Guglieri C Andrés A D Prieto C 2011 J. Phys. Chem. C 115 24092
17 Zhang L Li J Du Y Wang J Wei X Zhou J Cheng J Chu W Jiang Z Huang Y Yan C Zhang S Wu Z 2012 New J. Phys. 14 013033
18 Cheng Y Hao W Xu H Yu Y Wang T Chen R Zhang L Du Y Wang X L Dou S X 2012 ACS Appl. Mater. Interfaces 4 4470
19 Park C H Chadi D J 2005 Phys. Rev. Lett. 94 127204
20 Hao W Li J Xu H Wang J Wang T 2010 ACS Appl. Mater. Interfaces 2 2053
21 Wang Z H Geng D Y Guo S Hu W J Zhang Z D 2008 Appl. Phys. Lett. 92 242505
22 Manivannan A Dutta P Glaspell G Seehra M S 2006 J. Appl. Phys. 99 08M110
23 Sharma V K Varma G D 2007 J. Appl. Phys. 102 056105
24 Segura-Ruiz J Martinez-Criado G Chu M Geburt S Ronning C 2011 Nano Lett. 11 5322
25 Kohler T Armbruster T Libowitzky E 1997 J. Solid State Chem. 133 486
26 Du C L Gu Z B Lu M H Wang J Zhang S T Zhao J Cheng G X Heng H Chen Y F 2006 J. Appl. Phys. 99 123515
27 Bundesmann C Ashkenov N Schubert M Spemann D Butz T Kaidashev E M Lorenz M Grundmann M 2003 Appl. Phys. Lett. 83 1974
28 Alaria J Bouloudenine M Schmerber G Colis S Dinia A Turek P Bernard M 2006 J. Appl. Phys. 99 08M118
29 Samanta K Dussan S Katiyar R S Bhattacharya P 2007 Appl. Phys. Lett. 90 261903
30 Cong C J Liao L Liu Q Y Li J C Zhang K L 2006 Nanotechnology 17 1520
31 Han Y F Ramesh K Chen L Widjaja E Chilukoti S Chen F 2007 J. Phys. Chem. C 111 2830
32 Hu C C Wu Y T Chang K H 2008 Chem. Mater. 20 2890
33 Chen Y Hao W C Li W X Xu H Z Chen R Dou S X 2013 Chin. Phys. B 22 107501
34 Nogués J Sort J Langlais V Skumryev V Suriñach S Muñoz J S Baró M D 2005 Phys. Rep. 422 65
35 West K Nam D Lu J Bassim N Picard Y Stroud R Wolf S 2010 J. Appl. Phys. 107 113915
36 Sun Y Cong J Z Chai Y S Yan L Q Zhao Y L Wang S G Ning W Zhang Y H 2013 Appl. Phys. Lett. 102 172406
37 Martínez B Sandiumenge F Balcells L Arbiol J Sibieude F Monty C 2005 Phys. Rev. B 72 165202
38 Nagamiya T Yosida K Kubo R 1955 Adv. Phys. 4 1
39 Park J K Lee K W Kweon H Lee C E 2011 Appl. Phys. Lett. 98 102502