In GaN:Mg, the MgGa acceptor is compensated extensively by the formation of nitrogen vacancies (VN) and Mg interstitials (Mgi). However, we show that such compensation can be overcome by forming two kinds of Mg-rich complexes: one that contains VN and the other that contains only MgGa and Mgi. Such complexing not only neutralizes VN and Mgi but also forms better complex acceptors that have lower formation energies and smaller hole localization energies than isolated MgGa. Our results help explain the different doping behaviors in samples grown by different methods.

The realization of p-type conductivity in GaN:Mg has enabled the application of GaN-based electronic and optical devices. However, such industrial success is achieved under the shadow of a poor understanding of the underlying mechanism, despite decades of academic research. The p-type conductivity of GaN can be achieved only by heavily doping with Mg, with concentrations of Mg above 1019 cm−3 in metalorganic chemical vapor deposition (MOCVD) grown samples,1 for example. According to Hall-effect measurements of MOCVD GaN:Mg layers by Kaufmann et al.2 the hole concentration reaches a maximum value of 6 × 1017 cm−3 at an Mg concentration of 2 × 1019 cm−3, then decreases at higher Mg concentrations. The photoluminescence (PL) bands in undoped and Mg-doped GaN can help clarify the electronic properties, yet their attribution is still under debate. The ultraviolet luminescence (UVL) band is the dominant PL band in conductive n-type undoped and Mg-doped GaN grown by different techniques, as well as in conductive p-type Mg-doped GaN grown by molecular beam epitaxy (MBE). The blue luminescence (BL) band is observed only in GaN:Mg samples grown by hydride vapor phase epitaxy (HVPE) or MOCVD when the concentration of Mg exceeds 1019 cm−3.1 Kaufmann et al.3 connected the emergence of the room temperature (RT) PL band that peaks around 2.8 eV with the maximum value of the hole density reached as a function of Mg concentration for Mg concentration of (1–2) × 1019 cm−3 in MOCVD-grown GaN and presented clear evidence for the distant donor–acceptor (D–A) pair character of the 2.8 eV band. The reasons for the absence of p-type conductivity under moderate Mg doping, the onset of p-type conductivity under heavy Mg doping, and the subsequent drop-off in conductivity after reaching a maximum value of the hole concentration, as well as the origin of the UVL and BL in GaN:Mg, have been debated intensively for a few decades now, without resolution so far.

To date, the attribution of the electronic and optical properties of GaN:Mg to particular defect states remains unclear. Several models have been proposed. Apart from the commonly calculated deep levels of acceptors in wide bandgap semiconductors that localize a hole at single anion ligand atoms with a Jahn-Teller type symmetry breaking, Lany and Zunger4 found an effective-mass-like delocalized state of MgGa acceptors in GaN, which, in contrast, has a symmetric configuration. Despite the large geometry distortion in the localized state, this dual nature has a small energy difference of 0.03 eV, according to their calculation employing a Koopmans-corrected density functional theory (DFT) method. They predicted that the shallow state would dominate in n-type samples, whereas in less compensated p-type samples, the deep state would dominate. Lyons et al.,5 with hybrid DFT calculations, computed a relatively shallow (0/−) thermodynamic transition level for MgGa (260 meV), which they found to exhibit key features of a deep acceptor: the hole is localized on an N atom neighboring the Mg impurity, inducing a large local lattice distortion and giving rise to broad blue luminescence. Therefore, they attributed the BL to MgGa and the UVL to MgGa-Hi. They found no evidence for the existence of other metastable configurations, in contrast to Lany and Zunger. Using hybrid DFT calculations, however, Sun et al.6 found MgGa in GaN has a configuration characterized by highly anisotropic polaron localization that exhibits both effective-mass-like and noneffective-mass-like characters, similar to the result of Lany and Zunger. Their calculations were reproduced by Demchenko et al.,7 who attributed the UVL band to the anisotropically delocalized acceptor state. In earlier work, using the hybrid quantum mechanical (QM) and molecular mechanical (MM) embedded cluster method,8 we found that in lightly doped GaN:Mg, isolated MgGa strongly trap holes and, therefore, do not contribute to p-type conduction.

Whether the model involving only isolated MgGa defects, which may have a dual nature, is sufficient to account for the observed p-type conductivity and different PL bands or whether more complicated defects (complexes) have to be included has also drawn controversy in experimental studies. Monemar et al.9 studied bound-exciton spectra in samples for Mg concentrations up to [Mg] ≈ 1019 cm−3 and found evidence for two Mg-related acceptors in GaN. The unstable UVL emission peaking at 3.27 eV upon p-type activation was thus attributed to one of the two acceptors, which is unstable in p-GaN. Later, Monemar et al.10 attributed these two acceptors to isolated MgGa and a deeper acceptor that arises due to the interaction between basal plane stacking faults and MgGa acceptors. The dual nature model of Lany and Zunger was supported by Callsen et al.11 from the optical signature of Mg-doped GaN through the determination of the binding energies of two Mg acceptor states, while their sample was lightly Mg-doped (8 × 1017 cm−3). Singh et al.12 probed the lattice location of Mg in doubly doped GaN (Mg):Eu and linked two distinct Eu–Mg defect configurations with Lany and Zunger’s model. Electron paramagnetic spin resonance (EPR) and optically detected magnetic resonance (ODMR) measurements, along with PL, have also been performed to investigate acceptors in GaN:Mg. Glaser et al.13 performed ODMR and found a highly anisotropic g-tensor associated with the 3.27 eV PL emission from lightly Mg-doped samples predicted for effective-mass Mg acceptors in GaN. Later, (basal-plane) strain was highlighted as playing a crucial role in influencing the detailed nature of the acceptor states: the shallow effective-mass-like acceptor and deeper acceptor states are attributed to simple substitutional Mg at Ga sites in unstrained and strained regions, respectively.14,15

It is clear from the above-mentioned discussion that there is no single model that can satisfactorily explain the evolution of the p-type conductivity or the observed PL signals upon Mg doping. The electrical activation of Mg as a p-type dopant requires its incorporation into substitutional Ga as a single acceptor, while donor type defects such as N vacancies, Mg interstitials, or impurities such as hydrogen and silicon are likely to compensate, making MgGa a “vulnerable” acceptor dopant (giving rise to the observed doping limit16).

The nitrogen vacancy VN, as an intrinsic defect, has long been proposed as a major compensating center in GaN:Mg. By using room temperature Hall-effect measurements as a function of the Mg concentration, Kaufmann et al.2 found evidence for self-compensation, i.e., dopant-driven compensation of the Mg acceptor after reaching the maximum hole concentration by intrinsic donor defects rather than donor impurities, for which they developed a self-compensation model involving nitrogen vacancies. For heavily doped materials, they favored the VN-MgGa pair model. Applying positron annihilation spectroscopy, Hautakangas et al.17 recorded a positron lifetime of 180 ps in MOCVD grown Mg-doped GaN and suggested that neutral or negative complexes involving the monovacancy VN, in particular the VN-MgGa complex, were present in semi-insulating and p-type GaN. Interestingly, they observed no vacancies in MBE grown GaN:Mg layers. By performing Hall-effect, secondary-ion-mass spectroscopy (SIMS), and Rutherford backscattering spectrometry (RBS)-channeling measurements, Obata et al.18 investigated the lattice properties of heavily Mg-doped GaN and found that the lattice relaxation of nitrogen atoms is much higher than for Ga atoms, suggesting that the inverted n-type conductivity is caused by self-compensation effects induced by nitrogen vacancies. Using DFT calculations, Myers et al.19 indicated that VN compensates Mg acceptors extensively at higher doping levels. By employing a hybrid functional, Yan et al.20 found that both MgGa-VN and VN act as compensating centers in p-type GaN. Lee et al.21 studied three Mg-H-VN complexes in GaN:Mg and found that the inclusion of VN in many of the passivated complexes is necessary, given their abundance and energetic favorability in GaN:Mg. Using the QM/MM embedded cluster method,8 we previously determined that the nitrogen vacancy is a shallow compensating center in GaN doped with divalent metals, and consequently, isolated MgGa alone cannot account for the p-type conductivity in GaN:Mg. We calculated the optical level associated with VN at 44 meV below the conduction band minimum (CBM), to which we attributed the observed 3.466 eV UVL peak.

Apart from the nitrogen vacancy, there is evidence that the Mg interstitial (Mgi) is also a major compensating center in GaN:Mg. Early electronic structure calculations by Reboredo and Pantelides22 showed that Mgi and its complexes play a significant role in controlling the p-type doping process; they found that compensation of MgGa occurs mainly through the formation of Mg substitutional-interstitial pairs, among others, while H impurities can passivate them all. More recently, low formation energies for Mgi when the Fermi level is located below midgap have been computed.23 Miceli and Pasquarello,24 using hybrid DFT, found that the Mgi shows a noticeably lower free energy of formation than the MgGa in p-type conditions and, together with nitrogen vacancies, will act as a compensating center for donors. Recently, by implanting radioactive 27Mg into GaN, Wahl et al.25 gave direct experimental evidence for Mgi near octahedral sites and showed a significant concentration of Mgi in p-type GaN. Later, Wahl et al. elucidated the amphoteric nature of Mg, i.e., the concurrent occupation of substitutional Ga and interstitial sites. They concluded that challenges toward high electrical activation of implanted Mg are not related to a lack of substitutional incorporation.26 Along with other divalent metals, Be, for example, has also been investigated as an alternative p-type dopant; however, self-compensation induced by interstitials is likely to be considerably more pronounced for Be than for Mg, which would explain why no successful synthesis of p-type GaN:Be has been reported as yet.27 

To achieve p-type conductivity, the compensation effects of VN and Mgi induced inevitably by heavy Mg doping must be controlled. In this paper, we study complexes comprising the acceptor MgGa and the donors VN and Mgi. We show that VN as a triple donor in p-GaN must be passivated; MgGa as the only acceptor-type defect has to contribute to p-type conductivity, possibly by forming acceptor complexes; and Mgi as a double donor needs to be passivated in p-type GaN and can account for the doping threshold. We find that a large complex with Mg3N2 stoichiometry passivates VN and Mgi, and with one more MgGa as an acceptor— improves the p-type behavior of the dopant, a trend we postulate may continue for even larger clusters, although the small clusters studied here cannot explain p-type doping by themselves. We conclude that there could be two different acceptors: Mg-rich complexes with and without an N vacancy, which can explain the different doping behavior in GaN:Mg samples— grown by different methods.

In this work, electronic properties of defects and their complexes in wurtzite GaN:Mg are obtained following a hybrid QM/MM embedded cluster approach, which has been developed to describe localized states in ionic solids where charged and strong polar centers interact via long-range Coulomb and short-range Pauli repulsion forces.28,29 This method splits extended systems into the inner region, containing the centrally located defect, described using molecular QM theories, and its surroundings, which are only lightly perturbed by the defect, modeled with MM techniques based on interatomic potentials. Our choice of the QM methodology is DFT; the MM simulations employ polarizable-shell model interatomic potentials. The interface between the inner and outer regions employs cation-centered semi-local pseudopotentials to provide an appropriate embedding potential acting on the electronic subsystem of the defect and to contain the electrons within the inner region, effectively preventing them from spilling over the positively charged, attracting cations residing in the outer MM region. The outer part of the MM region is held fixed at the pre-optimized geometry,8,28 and any polarization here is treated a posteriori. The entire cluster is surrounded by point charges, whose values are fitted to reproduce the Madelung potential of the infinite crystal within the inner QM region along with its immediate MM environment, all contained within an “active sphere” of a user-defined radius (typically 15–20 Å). Using the embedded cluster approach with a suitably terminated QM cluster (i.e., with cations as terminating ions and an appropriate embedding potential), we can model the well localized valence band maximum (VBM) states. Therefore, by calculating the ionization energy of an electron in the bulk system, we can probe the VBM and determine its position relative to the vacuum. We then shift the CBM at an energy equal to the bandgap (3.5 eV) above the VBM relative to the vacuum. The advantage of our approach is the comprehensive account of short- and long-range polarization effects due to the presence of charged defects, the lack of periodic image interactions, the true infinitely dilute limit when reporting defect energetics, and an unambiguous definition of the vacuum level, facilitating the calculation of ionization energies with an absolute reference.

The method has been implemented in the ChemShell package,28,29,31 employing GAMESS-UK32 for the QM single-point energy and gradient calculations and GULP33 for the MM calculations. The QM region of 6.8 Å radius (116 atoms of wurtzite GaN) is embedded in an MM region of radius 30 Å. In our QM calculations, we use (i) the hybrid exchange and correlation density functional B97-2,34 which is similar to those commonly used in plane-wave supercell calculations (21% exact exchange compared with 25% for PBE035 or HSE0636), (ii) the SBKJC small-core pseudopotentials on Ga37 within the cluster and large-core refitted pseudopotentials8,30 on the cations forming the interface, and (iii) the atom-centered Gaussian basis set of def2-TZVP form on N38 and a matching quality SBKJC basis on Ga.37 For comparison, we use a second hybrid exchange and correlation density functional, BB1K, employing 42% exact exchange39 and fitted to reproduce kinetic barriers and thermochemical data, which removes a significant portion of the self-interaction error in exchange and correlation, thus giving a more accurate description of electron and hole localization and consequently deeper transition levels with respect to the VBM.40 Other computational details can be found in our earlier publication.28,41 We have applied this method previously to treat defects in AgCl,42 ZnO,43,44 and in earlier studies of GaN.8,40,45,46

We first show the formation energies under anion-rich conditions of the three important defects in p-GaN:Mg, as discussed earlier: VN, MgGa, and Mgi, in Figs. 1(a) and 1(b). We note that VN has been discussed in detail in our previous work.8,40 The formation energies of MgGa are reproduced from Ref. 8, but here we include a +/0 transition at 0.37 (0.84) eV above the VBM, determined using the B97-2 (BB1K) functional. Mgi stabilizes at an octahedral interstitial site in the hexagonal channel of wurtzite GaN, similar to Gai.40 Mgi is a donor-like defect for Fermi level values across the bandgap, stabilizing in the +2 charge state, and the formation energy at the VBM is −0.59 (−2.68) eV, computed using the B97-2 (BB1K) functional. For Fermi levels in the lower part of the bandgap, the formation energies of VN and Mgi are negative and lower than those of MgGa, implying spontaneous formation and a compensating nature for both. The low formation energies of VN and Mgi in p-GaN have also been reported in studies employing supercell calculations.20,23,24 Reshchikov et al. performed hybrid DFT calculations [using a modified Heyd–Scuseria–Ernzerhof (HSE) functional36 with an exact exchange fraction of 0.312] and found low formation energies of VN and Mgi near the VBM and the 0/− transition level of MgGa at 0.326 eV above the VBM.23 Miceli and Pasquarello performed similar hybrid DFT calculations (HSE functional with 0.31 exact exchange) and found similar low formation energies of VN and Mgi near the VBM and the 0/− transition level of MgGa at 0.38 eV above the VBM.24 Our results are consistent with the trend of the MgGa thermodynamic transition levels moving deeper into the gap with charged defects treated by our method compared with supercell calculations and with an increasing exact exchange fraction included in hybrid functionals, as detailed in Ref. 40. We thus confirm the compensating effects of VN and Mgi, a result largely independent of the functional or model used.47,48 Similar compensating effects have been found in GaN:Be.49 

FIG. 1.

(a) B97-2 and (b) BB1K calculated—formation energies of VN (black line), MgGa (red line), and Mgi (blue line) as a function of Fermi energy above the VBM. (c) and (e) B97-2 and (d) and (f) BB1K calculated—formation energies and binding energies of the complexes MgGa-VN (black line) and MgGa-Mgi (red line). Positive binding energies indicate favorable binding. Anion-rich conditions are assumed in (a)–(f). Structures of (g) MgGa-VN in a +2 charge state and (h) MgGa-Mgi in a +1 charge state. The blue, yellow, and red balls denote the Ga, N, and Mg atoms, respectively.

FIG. 1.

(a) B97-2 and (b) BB1K calculated—formation energies of VN (black line), MgGa (red line), and Mgi (blue line) as a function of Fermi energy above the VBM. (c) and (e) B97-2 and (d) and (f) BB1K calculated—formation energies and binding energies of the complexes MgGa-VN (black line) and MgGa-Mgi (red line). Positive binding energies indicate favorable binding. Anion-rich conditions are assumed in (a)–(f). Structures of (g) MgGa-VN in a +2 charge state and (h) MgGa-Mgi in a +1 charge state. The blue, yellow, and red balls denote the Ga, N, and Mg atoms, respectively.

Close modal

As triple- and double-donor defects in p-GaN, respectively, we expect VN and Mgi to easily bind the acceptor MgGa and form complexes MgGa-VN and MgGa-Mgi. We calculate three configurations of MgGa-VN, where the missing N, relative to the MgGa, is on either an axial site, a basal-plane site, or a next-nearest neighbor site. We find that these configurations all have very similar formation energies, and we present the results only for the axial site configurations.59 MgGa-Mgi is composed of Mgi with a neighboring Ga replaced by Mg. Figures 1(c) and 1(d) show the formation energies of these two complexes under anion-rich conditions (under anion-poor conditions, the formation energies will be 1.59 and 2.04 eV lower, respectively). They are donor-like in the p-type region (i.e., Fermi level values in the lower half of the gap) and have negative formation energies close to the VBM, so they will act as compensating centers. Comparing Figs. 1(a)1(d), we see that binding a MgGa results in VN and Mgi centers with slightly higher formation energies but, crucially, lower positive charge states in the p-type region and, consequently, less severe compensating centers as complexes than as isolated species. Note that the positive binding energies in Figs. 1(e) and 1(f) of these two complexes imply that they are bound species.45 Structures of MgGa-VN in the +2 charge state and MgGa-Mgi in the +1 charge state are shown in Figs. 1(g) and 1(h), respectively. The substitutional MgGa, as an isolated defect, tends to have associated with it a localized hole on a nearest-neighbor N ion in its neutral charge state and a second localized hole on another nearest neighbor N ion when the Fermi level is very close to the VBM (the defect is then in the + state). The energy needed to dissociate the localized hole is quite large, 1.404 eV as computed with the BB1K functional; therefore, the activation of MgGa and subsequent p-type conductivity are difficult to achieve. However, through binding a donor defect, the hole localized by MgGa recombines with a donor electron. Note that in both complexes, MgGa is in its state of −1 throughout the bandgap, except in MgGa-Mgi when the Fermi level is less than 0.07 (0.012) eV above the VBM for the B97-2 (BB1K) functional. Therefore, in p-GaN, these complexes will act as donors, but the MgGa-Mgi complex can release a trapped hole with an energy of 0.07 or 0.012 eV. However, crucially, both complexes can serve as nucleation centers for larger complexes, which may act as acceptors, as we show below.

The MgGa-VN complex has been studied several times previously. However, the MgGa-Mgi complex is much less well studied22 since the role of Mgi has only recently been appreciated. We compare our calculated formation energies of both complexes in the neutral charge state only with those determined using supercells since the periodic image interactions are particularly problematic for charged defects. For neutral MgGa-VN, we determine a formation energy of 2.32 (2.55) eV in N-rich conditions and 0.73 (0.96 eV) in N-poor conditions using the B97-2 (BB1K) functional, compared with supercell-method-calculated values of 2.8 eV (N-rich) and 2 eV (N-poor),20 2.1 eV (N-rich) and 1.25 (N-poor),24 determined with HSE, and 0.13 eV (N-poor),19 and 1.6 eV (N-poor)50 determined in earlier DFT studies. The compensating effects of MgGa and MgGa-VN have been investigated using various measurement techniques.2,17,18 Although VN are very difficult to detect experimentally due to either the small size of the vacancy, the high concentrations of other defects, or the requirements for negative and neutral charge states, depending on the experimental method,51 Hautakangas et al.17 proposed MgGa-VN complexes as native defects in Mg-doped GaN using positron annihilation spectroscopy (PAS). They attributed their measured 180 ps positron lifetime to VN or related complexes. As the isolated VN will not act as a positron trap because of its positive charge, they suggested that neutral or negative complexes involving VN are present in GaN:Mg. From computed core-electron momentum distributions, they identified MgGa-VN as the source of the observed positron lifetime. MgGa-VN is not neutral or negative in p-type GaN, as can be seen from our calculations and those of others referenced earlier. MgGa-VN is in the charge state of +2 until the +2/0 transition at 1.62 (2.13) eV above the VBM in Fig. 1(c) [and Fig. 1(d)]. Hybrid functional supercell calculations show this transition at ∼ 0.8 eV20,24,50 above the VBM. Like isolated VN, MgGa-VN has a positive charge state in p-type GaN and is unlikely to be a positron trap.

MgGa-VN and MgGa-Mgi, although less severe, are still compensating centers in p-type GaN, just like isolated VN and Mgi, respectively. Therefore, as double and single donor defects in p-GaN, we expect them to continue binding MgGa and to form complexes. Moreover, we expect such larger complexes to further reduce the compensating effect of such centers and help free holes from localization to MgGa, which may finally result in p-type conductivity in GaN:Mg.

Therefore, we calculate two Mg-rich complexes, 4MgGa-VN and 3MgGa-Mgi, where VN and Mgi are fully neutralized, and these two complexes are acceptors as a whole. 4MgGa-VN is composed of a N vacancy with its four nearest Ga all replaced by Mg, as shown in Fig. 2(a). In its closed-shell configuration of charge state −, the axial MgGa moves away from the vacant site by 43% (44%) of the equilibrium Ga–N bond length, while the other three basal MgGa move away by 19% (17%), as calculated using the B97-2 (BB1K) functional. When its charge state changes from − to 0 and then to +, MgGa moves farther away from the vacancy site by 4% when a hole is localized on a bonded N atom of MgGa. 3MgGa-Mgi is composed of an octahedral site Mgi with three of its six nearest Ga replaced by Mg, so we calculate one of six possible configurations, shown in Fig. 2(d). In Figs. 2(b), 2(c), 2(e), and 2(f), we show the formation energies in both conditions for these two complexes. They both have a lower formation energy under anion-poor conditions since the isolated MgGa, VN, and Mgi all favor anion-poor conditions. Comparing these two acceptor complexes with the isolated MgGa in Figs. 1(a) and 1(b), we find using the B97-2 functional (BB1K functional) that the 0/− transition level has been reduced from the isolated MgGa value of 0.863 (1.404) eV to 0.655 (0.997) and 0.690 (1.178) eV above the VBM, respectively, indicating that less energy is needed to dissociate a hole than that localized by isolated MgGa. Note that the MgGa-Mgi complex has a negative U effect, with a +/− transition level at 0.704 (1.237) eV determined using B97-2 (BB1K). Both complexes neutralize a major compensating center in p-GaN and become acceptors that have smaller hole dissociation energies compared to the isolated MgGa acceptor. Moreover, both complexes have lower formation energies in anion-poor conditions than isolated MgGa, indicating that they can form at higher concentrations than isolated MgGa in heavily doped GaN:Mg. We have calculated the equilibrium defect concentrations using the code SC-FERMI,52 as shown in Fig. 3.

FIG. 2.

Structures of (a) 4MgGa-VN and (d) 3MgGa-Mgi in −1 charge state. The blue, yellow, and red balls denote the Ga, N, and Mg atoms, respectively. Formation energies of (b) and (c) 4MgGa-VN and (d) and (e) 3MgGa-Mgi as a function of Fermi energy above the VBM calculated by B97-2 and BB1K functional, respectively. Anion-poor and anion-rich conditions are both given.

FIG. 2.

Structures of (a) 4MgGa-VN and (d) 3MgGa-Mgi in −1 charge state. The blue, yellow, and red balls denote the Ga, N, and Mg atoms, respectively. Formation energies of (b) and (c) 4MgGa-VN and (d) and (e) 3MgGa-Mgi as a function of Fermi energy above the VBM calculated by B97-2 and BB1K functional, respectively. Anion-poor and anion-rich conditions are both given.

Close modal
FIG. 3.

The calculated equilibrium concentrations of electrons (n), holes (p), and defects (complexes) studied in this work as a function of temperature. Results from both exchange and correlation density functionals are shown for (a) N-rich condition and (b) mid-way between the N-rich and N-poor conditions.

FIG. 3.

The calculated equilibrium concentrations of electrons (n), holes (p), and defects (complexes) studied in this work as a function of temperature. Results from both exchange and correlation density functionals are shown for (a) N-rich condition and (b) mid-way between the N-rich and N-poor conditions.

Close modal

Our calculations have, therefore, shown that the major compensating donor centers VN and Mgi in p-GaN:Mg can be neutralized by binding to multiple MgGa acceptors. Through forming a cluster with Mg3N2 stoichiometry and an additional MgGa, a large Mg-rich acceptor complex forms that is a better acceptor than isolated MgGa, with lower hole localization energies and lower formation energies in anion-poor conditions. Following this pattern, we can expect that even larger Mg-rich complexes can form in p-GaN and contribute to the p-type conductivity. Mg-rich pyramidal defects in p-GaN have been observed previously in experimental studies. For example, in MOVPE Mg-doped GaN, Vennéguès et al.53 found that these pyramidal defects are Mg rich and present in all studied films, independent of the doping level. Recently, maps of the electrostatic potential provided by off-axis electron holography have confirmed that the highest carrier concentration was achieved in the regions with the highest dopant concentration of 2 × 1020 cm−3, despite the presence of a high density of Mg-rich clusters revealed by correlating atom probe tomography.54 We propose that larger complex acceptors can have lower hole localization energies and, with the right size, form at least some of the shallow acceptors observed in conductive p-GaN. Kozodoy et al.55 investigated heavily Mg-doped GaN through variable-temperature Hall-effect measurements and found that the effective acceptor energy depth decreases from 190 to 112 meV and the compensation ratio increases as the dopant density is increased to the high values typically used in device applications. However, there is likely to be an optimum size of the Mg-rich complex since, in p-GaN, VN has a charge state of +3, and its formation energy will reduce drastically as the Fermi level moves closer to the VBM. As for Mgi, its formation energy is not as low as VN in p-GaN (though it will still decrease by twice the decrease in Fermi level as it shifts toward the VBM). However, heavy Mg-doping will increase the concentration of Mgi.25 Liliental-Weber et al.56 determined at atomic resolution of the Mg-rich hexagonal pyramids and truncated pyramids in GaN:Mg, to which they attributed the commonly observed decrease in acceptor concentration in the heavily doped material.

The BL band is probably due to a Mg-rich cluster that contains a VN since it is observed only in heavily doped GaN:Mg samples grown by HVPE or MOCVD,1 and Hautakangas et al.17 suggested that neutral or negative complexes involving isolated VN exist in semi-insulating and p-type MOCVD-grown GaN. In contrast, the BL band is not observed in p-type, Mg-doped GaN grown by MBE, and Hautakangas et al.17 observed no vacancies in MBE-grown GaN:Mg samples. This trend is consistent with that observed for the p-type conductivity; it is reported that the maximum hole concentration can be one order of magnitude higher in MBE-grown57 GaN:Mg than MOCVD-grown samples. As VN has a higher concentration in MOCVD-grown samples than in MBE-grown ones, we expect that the dominant compensating isolated donors in MOCVD samples will be VN and Mgi, and mostly Mgi in MBE-grown samples. Therefore, the overall compensating effect is more prominent in MOCVD samples, resulting in a different doping efficiency. The activation of p-type conductivity by annealing in MOCVD samples is probably due to the dissociation of large open volume defects into N monovacancy complexes, as can be seen in the PAS measurements of Ref. 17.

Several studies have proposed that the desired MgGa acceptor in as-grown MOCVD GaN:Mg is passivated by H donor-type defects by forming complexes of MgGa-Hi and that postgrowth annealing can dissociate those complexes and activate the isolated MgGa as acceptors.58 We argue that H, as a donor-type impurity, can passivate any form of acceptor in the sample,22 including the large complexes we propose here, so that H passivation is probably an additional process rather than a limiting factor in Mg doping.

In conclusion, we show that to achieve p-type conductivity in GaN:Mg, the major compensating centers VN and Mgi can be passivated by binding to MgGa centers. Such complexing can reduce the number of holes localized on isolated MgGa and form large Mg-rich acceptor complexes that have smaller hole localization energies and lower formation energies than isolated MgGa. There are two kinds of such large complexes, which differ as to whether they contain a VN. By comparison with experimental observations of the optical and electronic properties of GaN:Mg, we find that the complexes that contain a VN are likely to prevail in MOCVD-grown GaN but not in MBE-grown GaN, where only those containing Mgi and MgGa should be observed. Similarly, the BL band observed in MOCVD- and not MBE-grown GaN is attributed to VN-containing complexes. The requirement for annealing after growth in MOCVD- but not MBE-grown GaN to activate p-type conductivity is attributed to the need to dissociate large open volume defects from N monovacancy complexes. Future work will investigate the influence of complexes that contain more than one VN on p-type conductivity.

We are grateful to Professor Anthony Cheetham for many fruitful and enjoyable discussions relating to materials chemistry and physics. This paper continues our previous exciting collaboration with Aron Walsh, to whom we are indebted. The EPSRC is acknowledged for funding (Grant Nos. EP/K038419, EP/I03014X, and EP/K016288). Computational resources were provided by the ARCHER service through the Materials Chemistry High End Computing Consortium funded by EPSRC Grant No. EP/L000202, EP/R029431 and EP/X035859. GuangDong Basic and Applied Basic Research Foundation (Grant No. 2022A1515111156) is acknowledged.

The authors have no conflicts to disclose.

Zijuan Xie: Conceptualization (equal); Investigation (lead); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). John Buckeridge: Conceptualization (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Software (lead); Writing – review & editing (equal). C. Richard A. Catlow: Conceptualization (equal); Funding acquisition (equal); Methodology (equal); Supervision (lead); Writing – review & editing (equal). Anping Zhang: Data curation (equal); Formal analysis (equal); Validation (equal). Thomas W. Keal: Methodology (equal); Software (lead). Paul Sherwood: Software (equal). You Lu: Formal analysis (equal); Methodology (equal); Software (equal). Scott M. Woodley: Funding acquisition (equal); Methodology (equal); Resources (lead). Alexey A. Sokol: Conceptualization (equal); Methodology (equal).

The data that support the findings of this study are available from the corresponding authors upon reasonable request.

1.
M. A.
Reshchikov
,
P.
Ghimire
, and
D. O.
Demchenko
,
Phys. Rev. B
97
,
205204
(
2018
).
2.
U.
Kaufmann
,
P.
Schlotter
,
H.
Obloh
,
K.
Köhler
, and
M.
Maier
,
Phys. Rev. B
62
,
10867
(
2000
).
3.
U.
Kaufmann
,
M.
Kunzer
,
M.
Maier
,
H.
Obloh
,
A.
Ramakrishnan
,
B.
Santic
, and
P.
Schlotter
,
Appl. Phys. Lett.
72
,
1326
(
1998
).
4.
S.
Lany
and
A.
Zunger
,
Appl. Phys. Lett.
96
,
142114
(
2010
).
5.
J. L.
Lyons
,
A.
Janotti
, and
C. G.
Van De Walle
,
Phys. Rev. Lett.
108
,
156403
(
2012
).
6.
Y. Y.
Sun
,
T. A.
Abtew
,
P.
Zhang
, and
S. B.
Zhang
,
Phys. Rev. B
90
,
165301
(
2014
).
7.
D. O.
Demchenko
,
I. C.
Diallo
, and
M. A.
Reshchikov
,
Phys. Rev. B
97
,
205205
(
2018
).
8.
J.
Buckeridge
,
C. R. A.
Catlow
,
D. O.
Scanlon
,
T. W.
Keal
,
P.
Sherwood
,
M.
Miskufova
,
A.
Walsh
,
S. M.
Woodley
, and
A. A.
Sokol
,
Phys. Rev. Lett.
114
,
016405
(
2015
).
9.
B.
Monemar
,
P. P.
Paskov
,
G.
Pozina
,
C.
Hemmingsson
,
J. P.
Bergman
,
T.
Kawashima
,
H.
Amano
,
I.
Akasaki
,
T.
Paskova
,
S.
Figge
,
D.
Hommel
, and
A.
Usui
,
Phys. Rev. Lett.
102
,
235501
(
2009
).
10.
B.
Monemar
,
P. P.
Paskov
,
G.
Pozina
,
C.
Hemmingsson
,
J. P.
Bergman
,
S.
Khromov
,
V. N.
Izyumskaya
,
V.
Avrutin
,
X.
Li
,
H.
Morkoç
,
H.
Amano
,
M.
Iwaya
, and
I.
Akasaki
,
J. Appl. Phys.
115
,
053507
(
2014
).
11.
G.
Callsen
,
M. R.
Wagner
,
T.
Kure
,
J. S.
Reparaz
,
M.
Bügler
,
J.
Brunnmeier
,
C.
Nenstiel
,
A.
Hoffmann
,
M.
Hoffmann
,
J.
Tweedie
,
Z.
Bryan
,
S.
Aygun
,
R.
Kirste
,
R.
Collazo
, and
Z.
Sitar
,
Phys. Rev. B
86
,
075207
(
2012
).
12.
A. K.
Singh
,
K. P.
O’Donnell
,
P. R.
Edwards
,
K.
Lorenz
,
M. J.
Kappers
, and
M.
Boćkowski
,
Sci. Rep.
7
,
41982
(
2017
).
13.
E. R.
Glaser
,
M.
Murthy
,
J. A.
Freitas
,
D. F.
Storm
,
L.
Zhou
, and
D. J.
Smith
,
Phys. B
401–402
,
327
(
2007
).
14.
15.
M. E.
Zvanut
,
J.
Dashdorj
,
U. R.
Sunay
,
J. H.
Leach
, and
K.
Udwary
,
J. Appl. Phys.
120
,
135702
(
2016
).
16.
S. B.
Zhang
,
S. H.
Wei
, and
A.
Zunger
,
Phys. Rev. Lett.
84
,
1232
(
2000
).
17.
S.
Hautakangas
,
J.
Oila
,
M.
Alatalo
,
K.
Saarinen
,
L.
Liszkay
,
D.
Seghier
, and
H. P.
Gislason
,
Phys. Rev. Lett.
90
,
137402
(
2003
).
18.
T.
Obata
,
N.
Matsumura
,
K.
Ogiwara
,
H.
Hirayama
,
Y.
Aoyagi
, and
K.
Ishibashi
,
Phys. Status Solidi
3
,
1775
(
2006
).
19.
S. M.
Myers
,
A. F.
Wright
,
M.
Sanati
, and
S. K.
Estreicher
,
J. Appl. Phys.
99
,
113506
(
2006
).
20.
Q.
Yan
,
A.
Janotti
,
M.
Scheffler
, and
C. G.
Van De Walle
,
Appl. Phys. Lett.
100
,
142110
(
2012
).
21.
D.
Lee
,
B.
Mitchell
,
Y.
Fujiwara
, and
V.
Dierolf
,
Phys. Rev. Lett.
112
,
205501
(
2014
).
22.
F. A.
Reboredo
and
S. T.
Pantelides
,
Phys. Rev. Lett.
82
,
1887
(
1999
).
23.
M. A.
Reshchikov
,
D. O.
Demchenko
,
J. D.
McNamara
,
S.
Fernández-Garrido
, and
R.
Calarco
,
Phys. Rev. B
90
,
035207
(
2014
).
24.
G.
Miceli
and
A.
Pasquarello
,
Phys. Rev. B
93
,
165207
(
2016
).
25.
U.
Wahl
,
L. M.
Amorim
,
V.
Augustyns
,
A.
Costa
,
E.
David-Bosne
,
T. A. L.
Lima
,
G.
Lippertz
,
J. G.
Correia
,
M. R.
Da Silva
,
M. J.
Kappers
,
K.
Temst
,
A.
Vantomme
, and
L. M. C.
Pereira
,
Phys. Rev. Lett.
118
,
095501
(
2017
).
26.
U.
Wahl
,
J. G.
Correia
,
Â.R. G.
Costa
,
E.
David-Bosne
,
M. J.
Kappers
,
M. R.
da Silva
,
G.
Lippertz
,
T. A. L.
Lima
,
R.
Villarreal
,
A.
Vantomme
, and
L. M. C.
Pereira
,
Adv. Electron. Mater.
7
,
2100345
(
2021
).
27.
U.
Wahl
,
J. G.
Correia
,
A. R. G.
Costa
,
T. A. L.
Lima
,
J.
Moens
,
M. J.
Kappers
,
M. R.
Da Silva
,
L. M. C.
Pereira
, and
A.
Vantomme
,
Phys. Rev. B
105
,
184112
(
2022
).
28.
A. A.
Sokol
,
S. T.
Bromley
,
S. A.
French
,
C. R. A.
Catlow
, and
P.
Sherwood
,
Int. J. Quantum Chem.
99
,
695
(
2004
).
29.
P.
Sherwood
,
A. H.
de Vries
,
M. F.
Guest
,
G.
Schrecken-bach
,
C. A.
Catlow
,
S. A.
French
,
A. A.
Sokol
,
S. T.
Bromley
,
W.
Thiel
,
A. J.
Turner
,
S.
Billeter
,
F.
Terstegen
,
S.
Thiel
,
J.
Kendrick
,
S. C.
Rogers
,
J.
Casci
,
M.
Watson
,
F.
King
,
E.
Karlsen
,
M.
Sjøvoll
,
A.
Fahmi
,
A.
Schäfer
, and
C.
Lennartz
,
J. Mol. Struct.: THEOCHEM
632
,
1
(
2003
).
30.
M.
Miskufova
, Ph.D thesis,
University College London
,
2011
.
31.
S.
Metz
,
J.
Kästner
,
A. A.
Sokol
,
T. W.
Keal
, and
P.
Sherwood
,
Wiley Interdiscip. Rev.: Comput. Mol. Sci.
4
,
101
(
2014
).
32.
M. F.
Guest
,
I. J.
Bush
,
H. J.
Van Dam
,
P.
Sherwood
,
J. M. H.
Thomas
,
J. H.
Van Lenthe
,
R. W.
Havenith
, and
J.
Kendrick
,
Mol. Phys.
103
,
719
(
2005
).
33.
J. D.
Gale
,
J. Chem. Soc., Faraday Trans.
93
,
629
(
1997
).
34.
P. J.
Wilson
,
T. J.
Bradley
, and
D. J.
Tozer
,
J. Chem. Phys.
115
,
9233
(
2001
).
35.
C.
Adamo
and
V.
Barone
,
J. Chem. Phys.
110
,
6158
(
1999
).
36.
J.
Heyd
,
G. E.
Scuseria
, and
M.
Ernzerhof
,
J. Chem. Phys.
124
,
219906
(
2006
).
37.
W.
Stevens
,
M.
Krauss
,
H.
Basch
, and
P.
Jasien
,
Can. J. Chem.
70
,
612
(
1992
).
38.
F.
Weigend
and
R.
Ahlrichs
,
Phys. Chem. Chem. Phys.
7
,
3297
(
2005
).
39.
Y.
Zhao
,
B. J.
Lynch
, and
D. G.
Truhlar
,
J. Phys. Chem. A
108
,
2715
(
2004
).
40.
Z.
Xie
,
Y.
Sui
,
J.
Buckeridge
,
C. R. A.
Catlow
,
T. W.
Keal
,
P.
Sherwood
,
A.
Walsh
,
M. R.
Farrow
,
D. O.
Scanlon
,
S. M.
Woodley
, and
A. A.
Sokol
,
J. Phys. D. Appl. Phys.
52
,
335104
(
2019
).
41.
A.
Walsh
,
J.
Buckeridge
,
C. R. A.
Catlow
,
A. J.
Jackson
,
T. W.
Keal
,
M.
Miskufova
,
P.
Sherwood
,
S. A.
Shevlin
,
M. B.
Watkins
,
S. M.
Woodley
, and
A. A.
Sokol
,
Chem. Mater.
25
,
2924
(
2013
).
42.
D. J.
Wilson
,
A. A.
Sokol
,
S. A.
French
, and
C. R. A.
Catlow
, “
The Rôle of defects in photographic latent image formation
,” in
Materials Research Society Symposium Proceedings
(
MRS
,
2005
), Vol.
848
.
43.
A. A.
Sokol
,
S. A.
French
,
S. T.
Bromley
,
C. R. A.
Catlow
,
H. J. J.
van Dam
, and
P.
Sherwood
,
Faraday Discuss.
134
,
267
(
2007
).
44.
C. R. A.
Catlow
,
A. A.
Sokol
, and
A.
Walsh
,
Chem. Commun.
47
,
3386
(
2011
).
45.
Z.
Xie
,
Y.
Sui
,
J.
Buckeridge
,
A. A.
Sokol
,
T. W.
Keal
, and
A.
Walsh
,
Appl. Phys. Lett.
112
,
262104
(
2018
).
46.
Z.
Xie
,
Y.
Sui
,
J.
Buckeridge
,
C. R. A.
Catlow
,
T. W.
Keal
,
P.
Sherwood
,
A.
Walsh
,
D. O.
Scanlon
,
S. M.
Woodley
, and
A. A.
Sokol
,
Phys. Status Solidi A
214
,
1600445
(
2017
).
47.
D. O.
Demchenko
and
M. A.
Reshchikov
,
Phys. Rev. Lett.
115
,
029701
(
2015
).
48.
J.
Buckeridge
,
C. R. A.
Catlow
,
D. O.
Scanlon
,
T. W.
Keal
,
P.
Sherwood
,
M.
Miskufova
,
A.
Walsh
,
S. M.
Woodley
, and
A. A.
Sokol
,
Phys. Rev. Lett.
115
,
029702
(
2015
).
49.
D. O.
Demchenko
and
M. A.
Reshchikov
,
Appl. Phys. Lett.
118
,
142103
(
2021
).
50.
S.
Limpijumnong
and
C. G.
Van de Walle
,
Phys. Rev. B
69
,
035207
(
2004
).
51.
F.
Tuomisto
,
Proc. SPIE
8625
,
86250G
(
2013
).
52.
J.
Buckeridge
,
Comput. Phys. Commun.
244
,
329
(
2019
).
53.
P.
Vennéguès
,
M.
Benaissa
,
B.
Beaumont
,
E.
Feltin
,
P.
De Mierry
,
S.
Dalmasso
,
M.
Leroux
, and
P.
Gibart
,
Appl. Phys. Lett.
77
,
880
(
2000
).
54.
L.
Amichi
,
I.
Mouton
,
E.
Di Russo
,
V.
Boureau
,
F.
Barbier
,
A.
Dussaigne
,
A.
Grenier
,
P. H.
Jouneau
,
C.
Bougerol
, and
D.
Cooper
,
J. Appl. Phys.
127
,
065702
(
2020
).
55.
P.
Kozodoy
,
H.
Xing
,
S. P.
DenBaars
,
U. K.
Mishra
,
A.
Saxler
,
R.
Perrin
,
S.
Elhamri
, and
W. C.
Mitchel
,
J. Appl. Phys.
87
,
1832
(
2000
).
56.
Z.
Liliental-Weber
,
T.
Tomaszewicz
,
D.
Zakharov
,
J.
Jasinski
, and
M. A.
O’Keefe
,
Phys. Rev. Lett.
93
,
206102
(
2004
).
57.
I. P.
Smorchkova
,
E.
Haus
,
B.
Heying
,
P.
Kozodoy
,
P.
Fini
,
J. P.
Ibbetson
,
S.
Keller
,
S. P.
DenBaars
,
J. S.
Speck
, and
U. K.
Mishra
,
Appl. Phys. Lett.
76
,
718
(
2000
).
58.
S.
Nakamura
,
N.
Iwasa
,
M. S.
Masayuki Senoh
, and
T. M.
Takashi Mukai
,
Jpn. J. Appl. Phys.
31
,
1258
(
1992
).
59.

The formation energies of the configurations in the neutral and next-nearest neighbor cases; the corresponding +2/0 charge state are 2.55, 2.54 and 2.93 eV for the axial, basal-plane transition levels are 2.13, 2.05 and 2.22 eV above the VBM.