Introduction

Cadmium telluride (CdTe) and Cadmium zinc telluride (CdZnTe) are important materials for multiple applications in solar cells1,2,3, X-ray and gamma-ray room-temperature detectors4,5, electro-optic modulators6, and substrates for CdHgTe infra-red detector epitaxy7. In case of room-temperature semiconductor detectors these compounds overcome convention materials (Si, Ge, GaAs) by optimum conjunction of principal parameters - large atomic number of elements, large enough band gap energy, convenient electron mobility and attainable high resistivity and low enough trap density - necessary for successful detector manufacture8.

In spite of the high interest and expectation on CdTe/CdZnTe-detectors, till nowadays they have been hardly commercialized due to high expenses for fabrication high quality crystals. A CdTe crystal with a high resistivity is required for a good detector in order to reduce the dark current. Moreover a large mobility-lifetime product is desired to ensure that carriers generated by radiation in the whole detector volume can be collected by electrodes. Lattice defects form energy levels in the band gap and affect thereby the resistivity and the mobility lifetime product9. CdTe can be alloyed with Zn forming CdZnTe ternary compound. Alloying of CdTe with Zn strengthens the lattice and increases the band gap and thereby also maximum achievable resistivity10.

The electric properties are affected by lattice defects forming energy levels in the band gap9. The determination of the properties of native defects and their complexes is particularly important to understand their role in crystal growth and compensating of extrinsic defects4. In spite of extensive investigations the defects attached to distinct levels have in most cases not been completely identified. High purity CdTe/CdZnTe crystals are usually grown under Te-rich conditions because the Cd pressure required for Cd-rich conditions is as high as several atmospheres11. The CdTe/CdZnTe crystals grown under Te-rich conditions are typically characterized by low resistivity and p-type conductivity4 due to Cd vacancies which are divalent acceptors with two ionization levels and located below the mid-gap level4,12. To achieve high resistivity acceptors and residual impurities must be electrically compensated usually by doping using group III (Al, In, Ga) or group VIII (Cl) donors4,12.

Positron annihilation spectroscopy (PAS) including Positron lifetime (LT) and Coincidence Doppler broadening (CDB) spectroscopy was usually used for the study of the structure of point defects in CdTe/CdZnTe in the past13,14,15,16,17,18, but the link between defects and controlled annealing treatment has never been studied before.

This work reports on an investigation of the influence of defined annealing in Cd or Te vapour on the defect structure of Cl-doped CdTe and Ge-doped semi-insulating single crystals. The defect structure is characterized by PAS before and after each annealing step. The results are compared with galvanomagnetic Hall effect measurements in the Van der Pauw configuration19.

The paper is organized as follows. In Section Experimental we introduce samples, show the impurity content and summarize technical details of experiments. In Section Theory, we introduce methods for lifetime calculations and discuss their limits. Section Results and Discussion presents galvanomagnetic, positron back-diffusion measurements and theoretical calculations of lifetimes of positrons captured in different point defects. The subsection Lifetime spectroscopy reports the results of temperature dependent positron lifetime spectroscopy and their analysis based on three-state simple trapping model. This subsection is divided to the three parts describing for the undoped CdTe, Cl-doped CdTe and Ge-doped CdZnTe. The next subsection contains the results of CDB spectroscopy and their interpretation.

Experimental

The Cl-doped CdTe (CdTe:Cl) and Ge-doped , (CdZnTe:Ge) single crystals were grown by the vertical gradient freeze method20 (using 6N purity source elements). The annealing was done in the two zone furnace at 700 °C while the Cd or Te source temperature was 600 °C, which corresponds to Cd or Te pressure of 111 or 7.7 mbar, respectively. After annealing, the samples were cooled down at a cooling rate of 1 °C min−1. Nominally undoped CdTe (u-CdTe) single crystal grown by the vertical gradient freeze method at fairly increased Cd overpressure 1.6 atm was used as a reference sample.

LT measurements were carried out using a digital spectrometer21 with time resolution of 145 ps. 22Na activity of 1 MBq deposited on a 2 μm thick Ti foil was used as a positron source. The LT measurements were performed at 295 K (room temperature - RT) and at 123 K (low temperature), respectively. In both temperatures typically, 107 positron annihilation events were collected in LT spectra, which were subsequently decomposed into exponential components. The contribution of positron annihilations in the source spot and in the covering foils were always subtracted. In addition, temperature-dependent LT measurements were carried out in the temperature range of 123–295 K at a heating rate of ~1 K hour−1. Here the statistics of events was collected at each temperature step.

CDB spectroscopy22 was employed for characterization of local chemical environment of vacancies. The CDB studies were performed at RT using the same positron source as in the LT measurements. A digital CDB spectrometer23 equipped with two HPGe detectors and characterized by the energy resolution of 0.9 keV at the annihilation line and the peak-to -background ratio was employed for the CDB studies. At least annihilations were collected in each two-dimensional CDB spectrum. Subsequently the CDB spectra were reduced into one-dimensional cuts representing the resolution function of the spectrometer and the Doppler broadened annihilation peak. Normalized Doppler broadened peaks were divided by the normalized peak for the u-CdTe reference sample. Hence, in this paper the CDB results are presented as ratio curves with respect to u-CdTe.

Positron back diffusion measurement was performed on a continuous magnetically guided slow positron beam with energy of incident positrons adjustable in the range from 0.05 to 30 keV. Doppler broadening of the annihilation photopeak was measured by a HPGe detector with the relative efficiency of 35% and the energy resolution of  keV at 511 KeV. Shape of Doppler broadened annihilation photopeak was characterized using the S (sharpness) parameter24. The dependence of of the S parameter on positron energy was fitted using the VEPFIT code25.

Glow Discharge Mass Spectrometry (GDMS) was used for the chemical analysis. The concentrations of elements exceeding for u-CdTe, CdTe:Cl and CdZnTe:Ge presented in Table 1 in units 1015 cm−3. A 30% error of GDMS analysis was declared.

Table 1 The concentrations of elements exceeding 1015 cm−3 for u-CdTe, CdTe:Cl and CdZnTe:Ge established by GDMS in units 1015 cm−3.

Theory

Positron properties were calculated using density functional theory (DFT) within so-called standard scheme26. In this approximation positron density is assumed to be everywhere vanishingly small and not affecting the bulk electron structure. At first electron density in the material is solved without the positron. Subsequently, the effective potential for positron is constructed as

where is the Coulomb potential produced by the charge distribution of electrons and nuclei and is the zero positron density limit of the electron-positron correlation potential26.

Positron wave functions were calculated by solution of a single particle Schrödinger equation

where is the energy eigenvalue for i-th positron state. In the present calculations we considered the positron ground state only.

The positron annihilation rate (i.e. the inverse of positron lifetime) is determined using the expression

where is the classical electron radius, c is the speed of light, and γ denotes the electron enhancement factor describing the pileup of electrons at the positron site26.

The electron-positrons correlation, i.e. the correlation potential and the enhancement factor γ, were treated by two approaches:

  1. i

    local density approximation (LDA) utilizing the parametrization by Boroński and Nieminen27 and taking into account incomplete positron screening28 using a high frequency dielectric constant of 29 and

  2. ii

    generalized gradient approach (GGA) within the approach introduced by Barbiellini et al.30.

The electron density was constructed by superposition of atomic electronic densities calculated by a relativistic atomic code31. This approach called atomic superposition (ATSUP)32 neglects the charge transfer, but it is computationally feasible and can be used even for very large supercells retaining full 3-D geometry of the problem. In the following text the approach employing LDA and GGA scheme with superimposed electron density is denoted ATSUP-LDA and ATSUP-GGA, respectively.

The ATSUP-LDA and ATSUP-GGA calculations of positron parameters were performed on supercells containing 2048 Cd and 2048 Te atoms. Defects were modelled by removing the corresponding number of atoms from the supercell. Integration over the Brillouin-zone described in ref. 33 was used in calculations of positron parameters for defects.

It is well known that LDA overestimates positron annihilation rates especially with d electrons30 and the lifetimes calculated within the LDA approach are often shorter than the values measured in experiment. This shortcoming is to some extend compensated when LDA approach is used with electron density constructed by ATSUP32,34. On the other hand, the GGA approach when used with ATSUP electron density gives often lifetimes which are longer than the experimental values35. Hence, the lifetimes calculated using the LDA and the GGA approach can be considered as a lower and an upper bound of the interval where the actual lifetime falls. The GGA scheme is more sensitive to details of electronic structure than LDA30 and the lifetimes calculated using the GGA approach are it the best agreement with experiment when used with a self-consistent electron density. For these reasons we performed also self-consistent electron density calculations for selected positron states. The self consistent valence electron density was calculated by the plane wave code VASP (Vienna ab-inito simulation package)36,37 using projector augmented wave (PAW) potentials38. The , Cd electrons and , Te electrons were considered as valence electrons in the VASP calculations. The calculations were performed using 216 atoms based supercells and k-point grids generated using the Monkhorst-Pack scheme39. The wave functions were expanded in a plane wave basis with the cut-of energy of 277 eV. The calculated CdTe lattice constant a = 6.60 Å is in a reasonable agreement with the experimental value of 6.48 Å. Ionic relaxations were not considered at the present stage of calculations. In construction of the positron potential the frozen core electron orbitals were added to the self-consistent valence electron density calculated by VASP. In the following text this approach is denoted VASP-GGA.

The momentum distribution of annihilating pairs was calculated employing the ATSUP-based approach described in refs 40 and 41. The electron-positron correlations were treated within the GGA approach. The contribution from the i–th atom and a shell characterized by principal (n) and orbital (l) quantum numbers is calculated by the formula

where denotes the number of electrons in the shell, is the spherical Bessel function and and denotes the electron and positron radial wave functions. The symbol stands for the state-dependent positron enhancement factor41. The momentum distribution of the annihilating electron-positron pairs is obtained by summing the partial contributions over all occupied atomic sites and corresponding electron shells. Core electrons localized in atomic shells are practically not affected by crystal bonding. Hence, the ATSUP-based approach describes well the high momentum part of the momentum distribution where the contribution of positrons annihilated by core electrons dominates. But since it neglects the charge transfer it fails to describe properly the low momentum part of the momentum distribution defined by positrons annihilated by low momentum valence electrons.

In order to mimic the effect of the finite resolution of the experimental setup, the theoretical momentum distribution curves were convoluted with a Gaussian with FWHM of . To highlight the high momentum part of momentum distributions where the contribution of core electrons dominates the calculated momentum distributions are presented as ratio curves related to a perfect CdTe crystal.

Results and Discussion

Galvanomagnetic measurements

The galvanomagnetic properties of as-grown u-CdTe, CdTe:Cl and CdZnTe:Ge and annealed CdTe:Cl and CdZnTe:Ge samples are summarized in Table 2. The properties of u-CdTe accord with its growth at Cd overpressure, where the VCd-related defects are suppressed and the electron density is afforded by residual noncompensated donors. The as-grown CdTe:Cl and CdZnTe:Ge samples exhibit rather high resistivity, which is typical for the growth of Te-rich materials, where excess donors dominating above shallow acceptors are compensated by VCd8,42.

Table 2 Summary of galvanomagnetic measurements: resistivity of samples, type of conductivity and concentration of free carriers.

Annealing in Cd vapour leads to a decrease in resistivity for both samples. CdTe:Cl exhibited n-type conductivity due to the suppressed VCd and dominating shallow donors ClTe. The free carrier concentration of ~1017 cm−3 therefore gives a lower limit for the concentration of Cl dopants in the sample43. Cd vapour annealing of CdZnTe:Ge yield p-type conductivity contrary to CdTe:Cl. Such a behaviour is well known in annealed undoped or weakly doped CdTe/CdZnTe44, where acceptor-like impurities are dominant after annealing. In our CdZnTe:Ge, the extrinsic shallow donor ((B) + (Al)) and shallow acceptor ((P) + (Au) + (Cu)) densities are similar, and weak excess acceptor contamination is assumed. Subsequent annealing in Te vapour produced p-type conductivity in both samples. This is in agreement with expected VCd formation.

Calculated positron lifetimes

Positron lifetimes, calculated using ATSUP-LDA, ATSUP-GGA and VASP-GGA approaches for various positron states in CdTe are listed in Table 3. Theoretical positron lifetimes available in literature are shown in Table 3 as well. The positron lifetimes calculated using the ATSUP-LDA and ATSUP-GGA schemes in the present work are in an excellent agreement with the lifetimes calculated in refs 15,45. One can see in Table 3 that positron lifetimes calculated using various approaches for the electron-positron correlation, i.e. the enhancement factor and the electron-positron correlation potential, differ. The positron lifetimes calculated using the LDA approach are always lower than those obtained by the GGA scheme. Moreover, calculated positron lifetimes vary depending whether superimposed or self-consistent electron density was used in the calculation. The bulk positron lifetime for CdTe calculated using the ATSUP-LDA approach is  ps, while the ATSUP-GGA scheme yielded higher bulk lifetime  ps. For comparison experimental lifetimes reported in literature for CdTe are listed in Table 4. The experimental bulk positron lifetimes for CdTe reported in literature fall into the range from 280 to 291 ps13,14,15,16,18,46,47. The VASP-GGA scheme which can be considered as the most precise approach used here yielded the CdTe bulk positron lifetime  ps which is comparable with the experimental values.

Table 3 Theoretical positron lifetimes for a perfect CdTe crystal and various defects calculated by various approaches.
Table 4 Experimental positron lifetimes for a perfect CdTe crystal and selected defects reported in literature.

In order to compare positron lifetimes for various defects with the experimental values we use ratios of the calculated lifetime to the bulk lifetime calculated by the same approach. These ratios are listed in Table 5. The differences between various theoretical schemes are to a large extend cancelled using ratios and the values exhibit relatively low sensitivity to the approach used for electron-positron correlation30,33.

Table 5 Theoretical τD/τB ratios for positrons trapped at various defects in CdTe calculated by various approaches.

While is believed to be positively charged in CdTe, is either neutral or negatively charged depending on the Fermi level position12. Hence, positrons are repelled by while represent trapping sites for positrons. This is illustrated in Fig. 1 presenting the positron density in the (001) plane calculated by the ATSUP-GGA approach. Figure 1a shows the positron density in a perfect CdTe crystal where positron is de-localized in the lattice. The positron density calculated for a CdTe crystal containing is plotted in Fig. 1b. Obviously the positron is localized in . Experimental evidence for positron trapping in has been reported by many authors14,15,16,18,46. The lifetimes of positrons trapped determined experimentally fall into the range from 320 to 325 ps which corresponds to falling into the interval 1.11–1.15, see Table 4. The ATSUP-LDA and ATSUP-GGA calculations for positrons trapped at resulted in , which is remarkably lower than the experimental values. On the other hand, the VASP-GGA approach yielded which is close to the values measured in experiment. This indicates that charge transfer which was neglected in ATSUP calculations leads to a slight decrease of electron density in and consequently an increase of the lifetime of trapped positrons.

Figure 1
figure 1

Positron density in the (001) plane calculated using the ATSUP-GGA approach (a) perfect CdTe crystal, (b) CdTe crystal containing VCd in 1, 1, 0 position, (c) CdTe crystal containing AC defect consisting of VCd and ClTe located in 1, 1, 0 and 0.25, 1.75, 0.25 position, respectively. The co-ordinates are expressed in the units of the CdTe lattice constant a. Positron density is given in the atomic units.

Positron lifetime was calculated also for associated with Te anti-site and with various impurities. One can see in Table 3 that causes shortening of positron lifetime. A donor associated with acceptor forms complex which is electrically neutral. Theoretical calculations revealed that positrons trapped at complexes exhibit practically the same lifetime as . Hence, these two kinds of defects cannot be distinguished by measurement of positron lifetime.

In Cl doped CdTe may couple with shallow donors forming negatively charged complexes called A-centres or neutral complexes. Results in Table 3 indicate that replacement of Te nearest neighbours of by Cl atoms increases the lifetime of trapped positron. This effect is caused by smaller size of Cl atoms compared to Te ones. Hence replacement of a Te nearest neighbour of by a Cl impurity increases the open volume of vacancy. This is illustrated in Fig. 1c which presents calculated positron density in the (001) plane for a positron trapped at AC defect. One can see in the figure that positron density in AC becomes asymmetric since it expanded towards the Cl nearest neighbour.

Lifetime spectroscopy

The results of LT measurements, i.e. lifetimes and relative intensities resolved in LT spectra for u-CdTe, CdTe:Cl and CdZnTe:Ge samples, are summarized in Table 6 for temperatures 295 and 123 K, respectively. The temperature dependences of the mean positron lifetime are presented in Fig. 2.

Table 6 Results of LT studies: lifetime τi and the relative intensity Ii of the exponential components resolved in the LT spectra evaluated at 295 K and 123 K.
Figure 2: Temperature dependence of the mean positron lifetime for CdTe:Cl and CdZnTe:Ge.
figure 2

The lines represent the fit using 3-STM.

Undoped CdTe

The u-CdTe crystal exhibits a single component LT spectrum (except of the source contribution) with lifetime of ~290 ps, which fits in the CdTe bulk lifetimes reported in literature, see Table 4. Hence, the concentration of positron traps in the u-CdTe crystal is very low (below the sensitivity threshold of LT spectroscopy) and virtually all positrons are annihilated in the de-localized state. Low temperature LT measurements of the u-CdTe crystal yielded practically the same results as the measurement at RT.

Results of positron back-diffusion measurement, i.e. dependence of the S parameter on the energy E of incident positrons in plotted in Fig. 3. The positron penetration depth increases with increasing energy of implanted positrons. Hence, at very low energies virtually all positrons are annihilated at the surface. With increasing energy positrons penetrate deeper into the sample and the fraction of positrons diffusing back to the surface gradually decreases. Finally at high energies virtually all positrons are annihilated in the bulk and the S parameter approaches a plateau. One can see in Fig. 3 that the curve exhibits a maximum at energy around 2 keV. This indicates that the sample contains a thin surface layer of native oxides most probably a mixture of , and CdO48,49. The curve for the u-CdTe sample was fitted by VEPFIT25 using a two layer model: (i) native oxide layer on the surface and (ii) CdTe bulk region. The model curve calculated by VEPFIT is plotted in Fig. 3 by a solid line and is obviously in good agreement with the experimental points. The thickness of the native oxide layer ≈30 nm was fitted.

Figure 3: The dependence of the S parameter on the energy E of incident positrons for undoped CdTe (full symbols) and CdTe:Cl sample in the as-grown state (open symbols).
figure 3

Solid lines show model curves calculated by VEPFIT.

The positron diffusion length determined in the CdTe region is  nm. This value is comparable with the positron diffusion lengths in defect-free semiconductors24. Hence, the u-CdTe crystal contains very low density of positron traps (except of the thin surface oxide layer) and almost all positrons are annihilated from the free state in accordance with the results of positron lifetime characterization. Assuming that the u-CdTe sample contains no positron traps one can calculate the positron diffusion coefficient for u-CdTe from the expression , where  ps is the bulk u-CdTe lifetime. This yields the positron diffusion coefficient for u-CdTe  cm2s−1 which is in very good agreement with the value  cm2s−1 reported for CdTe by Neretina et al.50.

The u-CdTe crystal exhibits very low density of dislocations which does not exceeds 109 m−2 as determined by etch pitting technique. We performed an additional independent test whether positrons in u-CdTe crystal are annihilated in the free state and the lifetime of ~290 ps measured in this sample can be indeed considered as the bulk lifetime. The u-CdTe sample was mechanically grinded by F600 SiC paste with grain size 9 μm. Plastic deformation caused by grinding introduced dislocations into the sample. The grinded u-CdTe sample exhibits two-component LT spectrum with lifetimes  ps and  ps and corresponding relative intensities % and %. Hence, it is clear that grinding introduced defects into the sample. The lifetime of the second component is comparable with the lifetime of positrons trapped at (see next section). This is typical for positrons trapped at dislocations24,51. Thus, the component with lifetime can be attributed to positrons trapped at dislocations introduced by grinding. The lifetime of the first component was shortened with respect to the as-grown u-CdTe sample. This is in accordance with the two-state simple trapping model24,52. When positrons are trapped at defects then the lifetime of the free positron component becomes shorter than that in the defect-free material and the quantity

equals to the positron bulk lifetime, i.e. 24,52. For the grinded u-CdTe Eq. (5) yielded ps which is in an excellent agreement with the lifetime measured in the as-grown u-CdTe. This gives strong evidence that the lifetime of ~290 ps measured in the as-grown u-CdTe comes indeed from annihilations of free positrons. If the lifetime measured in the as-grown u-CdTe would come from positrons trapped at some defects, then introduction of additional defects (here dislocations) by grinding would not cause any shortening of the lifetime .

CdTe:Cl

The LT spectra of the CdTe:Cl crystal generally consist of up to three components (except of the source contribution) corresponding to the three positron states: (i) free positrons (lifetime τ1), (ii) positrons trapped in shallow traps (lifetime , trapping rate , and (iii) positrons trapped in deep traps (lifetime , trapping rate . The kinetics of positron trapping for these defects is described by the three-state simple trapping model (3-STM)24. The rate equations for such system read24

with the initial conditions and . The symbols , and denote the probabilities that a positron is at time t de-localized in the free state or trapped in a shallow or deep trap, respectively. and stand for the positron trapping rates for the shallow and the deep traps, respectively, and is the de-trapping rate from the shallow traps. The lifetimes of positrons trapped at the shallow and the deep traps, i.e. inverse of the annihilation rates of trapped positrons, are denoted and , respectively. Solution of the differential equations (6) gives the decay spectrum of positrons

where

and

The LT spectrum is the number of positrons annihilated in various times and is obtained as a negative time derivative of the decay spectrum

Hence the lifetimes , , of the exponential components resolved in experimental LT spectra and their relative intensities , , can be interpreted within 3-STM using Eqs (8. Note that for a high concentration of shallow or deep positron traps the lifetime becomes extremely short and its intensity diminishes. In such case it is not possible to resolve the first component in LT spectrum due to limited resolution of LT spectrometer. This situation happened here in the case of low temperature LT spectra for CdTe:Cl. The short lived component could not be resolved in LT spectra due to very high density of shallow traps in this sample as will be shown in the following text.

The probability that a positron trapped in a shallow trap will escape by thermal excitation decreases with falling temperature. Hence the positron de-trapping rate from shallow traps depends on temperature53

where is the effective positron mass26, the Boltzmann constant, [c] the concentration of shallow traps, and the positron binding energy to the shallow traps. The temperature dependence of the mean lifetime was fitted using 3-STM with , , [c] and as fitting parameters. If the sample contains shallow traps, then the mean positron lifetime varies with temperature due to the temperature dependence of . Figure 2 shows that for CdTe:Cl decreases with temperature, which testifies to the presence of shallow traps in this sample. The temperature dependence of can be explained by the formation of series of attractive shallow Rydberg states connected with negatively-charged defects54,55. Since the positron binding energy of the shallow Rydberg states is rather small (typically ≤0.1 eV), they act as efficient positron traps at low temperatures only where the positron may not escape by thermal excitation54.

Hence, to describe LT spectra measured at RT 3-STM can be simplified by elimination of the shallow traps, i.e. by setting , in Eqs (6. This leads to the following rate equations

This simplified trapping model called two-state simple trapping model (2-STM) was used for analysis of the LT spectra measured at RT. The solution of 2-STM with the initial conditions and is two-component decay spectrum of positrons

where

The LT spectrum predicted by 2-STM is

Decomposition of LT spectrum of the as-grown CdTe:Cl measured at RT revealed two exponential components, see Table 6. The shorter component with lifetime ps represents a contribution of free positrons. Note that is shorter than the bulk positron lifetime for CdTe due to parallel positron annihilation and trapping in the deep traps, c.f. Eq. (14). The lifetime ps, , of the longer component is significantly higher than the positron lifetime for single vacancies in CdTe, see Table 3. Hence, we deduce that positrons in the as-grown CdTe:Cl are trapped in larger point defects with the open volume comparable with that of several vacancies. The positron trapping rate to these vacancy clusters can be calculated within 2-STM from Eqs. (14)

The concentration of vacancy clusters is directly proportional to the trapping rate determined by 2-STM, . The specific positron trapping rate for the vacancy clusters was estimated as , where is the specific positron trapping rate for neutral vacancies in semiconductors24 and is the average number of vacancies constituting the cluster. The latter value was estimated by comparison of the ratio determined experimentally with the theoretical calculations of clusters of the neutral complexes of various sizes in Table 5.

Extended investigations of CdTe:Cl with chlorine content from 100 to 3000 ppm performed by Krause-Rehberg et al.13 revealed two components in LT spectra measured at RT: (i) the shorter component, with a lifetime of 330(10) ps, , which was attributed to and (ii) a long-lived component with a lifetime of 450(15) ps, , which was assigned to clusters of neutral complexes. The concentration of both defects increased with Cl content, testifying that they are associated with chlorine. Our CdTe:Cl sample exhibits a smaller lifetime than 450 ps. This could be due to significantly lower Cl concentration leading to the formation of clusters of smaller size.

The analysis of the LT spectrum of the as-grown CdTe:Cl measured at 123 K revealed a new component with lifetime  ps, which is shorter than the bulk positron lifetime (290 ps) and dominates the spectrum with 94% intensity, see Table 4. Taking into account the temperature dependence of in Fig. 2, this component obviously comes from positrons trapped in the shallow Rydberg states associated with negatively-charged A-centres . Indeed, the positron binding energy  eV determined from fitting of the temperature dependence of falls into the range  eV expected for the Rydberg states52. Evidently, . The concentration obtained from fitting of LT data by 3-STM is presented in Table 7. From the comparison of with we can estimate the specific positron trapping rate for the shallow traps  s−1. This value falls into the expected range for calculated in ref. 54.

Table 7 The concentration of defects calculated from the LT data.

Note that Rydberg states are characterized by a lifetime which is close to the bulk lifetime54,55. Depending on the local electron density the lifetime of positrons confined at low temperatures in shallow traps can be either higher or lower than the bulk positron lifetime. The latter case has been recently reported for Cu in-diffused GaAs:Te crystals56. Low temperature LT studies of Cu in-diffused GaAs:Te revealed shallow traps characterized by a lifetime of 220 ps, which is about of 8 ps shorter than the GaAs bulk lifetime of 228 ps56. The shallow traps in Cu-indiffused GaAs:Te were identified as ions coupled with Ga vacancies . At RT positrons in the Cu-indiffused GaAs:Te sample are trapped at deep traps identified as complexes and characterized by a positron lifetime of 280 ps. At low temperatures positrons are confined predominantly in Rydberg states associated with ions and characterized by a positron lifetime of 220 ps, which is shorter than the bulk lifetime56. Thus, it seems that positron trapping in the CdTe:Cl sample is to some extent analogous to that in Cu-indiffused GaAs:Te: at RT positrons are trapped at deep traps (vacancy clusters) while at low temperatures positrons are confined at Rydber states associated with Coulomb field around negatively charged A-centers and characterized by a positrons lifetime  ps, which is about of  ps shorter than the CdTe bulk lifetime of  ps measured on the u-CdTe sample, see Table 4. We assume that the reduced lifetime of the positron in the Rydberg state comparing to the bulk lifetime may stem from a lattice relaxation and enhanced mass density in the vicinity of the vacancy, at the radius about 1 nm, where the positrons in Rydberg state mostly occur.

Annealing the CdTe:Cl sample in Cd vapour led to a significant reduction of the intensity of the long-lived defect component and shortening of its lifetime down to ps at RT. The ratio is close to the value calculated for or single complex The reduction of this component is clearly caused by the removal of by Cd annealing and the reduction in cluster size. Ensuing analysis within 2-STM proves the depression of the density of related defects, see Table 7. The Cd-annealed CdTe:Cl is characterized by very low resistivity and n-type conductivity, caused by shallow donors which in contrast to the as-grown sample are not compensated by shallow acceptors. In spite of significant depreciation the sample measured at low temperature again exhibits a remarkable contribution of , see Table 6. The temperature dependence of plotted in Fig. 2 is similar to the as-grown sample just shifted to lower values due to the suppressed contribution from neutral clusters. Since the lifetime of positrons localized in the shallow traps is lower than the CdTe bulk lifetime, decreases below the bulk lifetime value at low temperatures where almost all positrons are confined in the shallow traps. Referring to similar concentration of defects at RT and at low temperature, see Table 7, we deduce that in the Cd-annealed CdTe:Cl sample is associated with only.

Subsequent annealing of a CdTe:Cl sample in Te vapour leads to the appearance of a defect component with lifetime  ps, which corresponds to negatively-charged and neutral defects. Unlike the as-grown CdTe:Cl crystal, the complexes formed after Te-annealing are isolated and do not agglomerate into clusters. This could be caused by the fact that during Te-annealing the sample was kept at a lower temperature than during crystal growth, and the mobility of the complexes was therefore lower. The analysis of LT spectra measured at 123 K revealed that the positron trapping in dominates at low temperatures in a similar way as in the Cd-annealed crystal. The temperature dependence of presented in Fig. 2 again shows a drop with decreasing temperature in a similar manner as for the Cd-annealed sample.

Referring to the concentration of defects (Table 7), we can deduce that the defects observed at RT in the Te-annealed sample are associated with and at the total density . Similarly, the low temperature measurements allow us to establish , which yields . Consequently, the calculation of the total Cl density in -related complexes provides . This result is very close to the concentration of chlorine determined by GDMS . Since GDMS analysis is affected by up to 30% error, we consider this finding as an excellent confirmation of our assignment of defects.

Similar estimation can be done also for the as-grown CdTe:Cl sample where the concentration of clusters of was estimated in Table 7. Assuming that each cluster consists on average of 4 complexes (as deduced from the ratio the concentration of Cl contained in clusters is . At the same time the low temperature measurement of the same sample yielded . Hence, the total concentration of Cl in -related complexes in the as-grown CdTe:Cl is . Again this value is in very reasonable agreement with the total Cl density determined by GDMS.

The as-grown CdTe:Cl crystal was characterized also by positron back-diffusion measurement on the slow positron beam. The dependence of the S-parameter on the energy E of incident positrons for the as-grown CdTe:Cl crystal is plotted in Fig. 3. Obviously the curve for the as-grown CdTe:Cl sample exhibits similar features as the curve for the u-CdTe crystal, namely a maximum at low energies (≈2 keV) followed by a gradual decrease down to a plateau value at high energies where virtually all positrons are annihilated inside the CdTe:Cl region. However the plateau value at high energies for the as-grown CdTe:Cl sample is remarkably higher than that for the u-CdTe. The curve was fitted in the similar manner as the curve for u-CdTe sample, i.e. using a two-layer model consisting of a thin surface oxide layer and the bulk CdTe:Cl region. The model curve calculated by VEPFIT25 is plotted by solid line in Fig. 3 and describes the experimental points accurately. The thickness of the oxide layer of ≈20 nm obtained from fitting is comparable with that measured in the u-CdTe. The positron diffusion length in the CdTe:Cl region is  nm. This value is shorter than the positron diffusion length  nm determined for the u-CdTe. Shortened positron diffusion length and higher S parameters at high energies give clear evidence that the as-grown CdTe:Cl sample exhibits higher density of defects than the u-CdTe. The concentration of defects [c] in the as-grown CdTe:Cl sample can be estimated from shortening of the positron diffusion length using the expression57

where v is the specific positron trapping rate to defects. From LT studies we know that the as-grown CdTe:Cl sample contains clusters consisting on average of four . Assuming one gets which is in very reasonable agreement with the concentration of clusters determined by LT spectroscopy in Table 7.

CdZnTe:Ge

At RT the CdZnTe:Ge as-grown crystal exhibits a defect component with lifetime  ps falling into the range  ps commonly attributed to shallow acceptors or complexes of associated with donors, see Table 4. The lifetime at 123 K is practically the same as in RT, and remains constant apart from the statistical scattering in the whole temperature range of 123–295 K, see Fig. 2. A similar result was observed17 for undoped CdZnTe. The temperature independence of observed in CdZnTe:Ge crystal testifies that (i) the defects responsible for the lifetime in this sample are mostly electrically neutral, and (ii) unlike CdTe:Cl, the CdZnTe:Ge crystal does not contain shallow traps in detectable concentration.

Taking into account electrical neutrality of defects and positron lifetime comparable to the component with the lifetime can be attributed to positrons trapped at electrically neutral complexes which are likely formed in CdZnTe:Ge. Table 7 shows the concentration of complexes determined as , where is the trapping rate calculated within 2-STM by Eq. (16) and is the specific positron trapping rate for neutral vacancies in semiconductors24.

Annealing of CdZnTe:Ge crystal in Cd vapour led to the disappearance of the defect component with lifetime , and the LT spectrum became single-component with the lifetime  ps, which is close to the bulk lifetime of 292(1) ps measured in the u-CdTe, see Table 6. Hence, virtually all positrons in the Cd-annealed CdZnTe:Ge sample are annihilated in the de-localized state. This result proves that annihilation is due to Cd-annealing. Slightly higher value of the bulk positron for CdZnTe:Ge compared to that for u-CdTe could be caused by lower electron density in CdZnTe compared to that in CdTe. Temperature-dependent measurements did not reveal any temperature dependence of LT results for the Cd-annealed CdZnTe:Ge sample, see Fig. 2 and Table 6, which is in accordance with the assignment of the component to the free positrons.

The subsequent annealing of CdZnTe:Ge in Te vapour resulted in the restoration of the defect component with lifetime  ps, which shows the re-emergence of . The LTs obtained after annealing are very similar to those measured in the as-grown CdZnTe:Ge, including the temperature independence of in Fig. 2, which testifies to positron trapping in electrically neutral complexes.

CDB spectroscopy

The CDB spectroscopy was employed in order to obtain direct information about chemical environment of in CdTe:Cl and CdZnTe:Ge crystals. Figure 4 shows the momentum distribution of annihilating electron-positron pairs measured in the u-CdTe reference sample. The calculated momentum distribution for a perfect CdTe crystal is plotted in the figure as well. One can see in the figure that the momentum distribution calculated using the ATSUP-GGA approach agrees well with the experimental points in the high momentum range where the contribution of positrons annihilated by core electrons dominates. This is caused by the fact that core electrons in the inner shells retain their atomic character are almost not influenced by crystal bonding. On the other hand the ATSUP-based approach is not able to describe accurately the momentum distribution in the the low momentum region where the contribution of positrons annihilated by valence electrons becomes dominating. For this reason the calculated momentum distributions will be compared with the experimental data in the high momentum region only.

Figure 4
figure 4

The positron annihilation momentum density P(pz) measured by CDB in the undoped CdTe sample and calculated for a perfect CdTe crystal using the ATSUP-GGA approach.

Figure 5a shows the experimental CDB ratio curves related to the u-CdTe reference for the CdTe:Cl sample in the as-grown state and after annealing in Cd and Te vapour, respectively. The calculated CDB curves related to a perfect CdTe crystal for , and defects are plotted in Fig. 5b. From inspection of Fig. 5 it becomes clear that the CDB curve for CdTe:Cl annealed in Cd vapour is almost flat (except of statistical scattering) and close to unity in the whole momentum range testifying that the momentum distribution for the Cd-annealed CdTe:Cl is very similar to that in the u-CdTe reference. This confirms the picture that positron traps were removed by annealing in Cd vapour and virtually all positrons in the Cd-annealed sample are annihilated in the free state. On the other hand, the CDB ratio curves for the as-grown CdTe:Cl and the sample annealed in Te vapour clearly differ from unity and exhibit a distinct peak in the high momentum range at . Similar peak is observed in the CDB curves calculated for positrons trapped in and defects in Fig. 5b. This testifies to positron trapping in defects associated with chlorine in the as-grown and Te-annealed CdTe:Cl sample.

Figure 5
figure 5

The CDB ratio curves (related to defect-free u-CdTe) (a) experimental data for CdTe:Cl in the as-grown state and after annealing in Cd and Te vapour; (b) calculated ratio curves for VCd, AC and VCd2ClTe defects. The calculated curves are plotted by dashed lines in the low momentum range (pz < 15 × 10−3m0c) where the accuracy of the momentum distribution calculated by ASTUP is poor.

The CDB results for CdZnTe:Ge samples are presented in Fig. 6. The experimental CDB curves for the as-grown and Cd or Te-annealed CdZnTe:Ge are plotted in Fig. 6a while Fig. 6b shows the CDB ratio curves calculated for positrons trapped at and defects. The CDB ratio curve for Cd-annealed CdZnTe:Ge sample is again almost flat and close to unity indicating that the momentum distribution in this sample is similar to that in the u-CdTe reference. A slight enhancement in the momentum range is most probably due to positrons annihilated by electrons belonging to Zn. Annealing in Te vapour modified the momentum distribution of CdZnTe:Ge sample. Shape of the ratio curve for the Te-annealed sample is very similar to the curves calculated for positrons trapped at and defects. Hence, in accordance with the results of LT spectroscopy the CDB data in Fig. 6 indicate that positrons in CdZnTe:Ge annealed in Te vapour are trapped at defects containing . Very similar CDB curve was observed also in the as-grown sample. However, CDB spectroscopy does not allow to discern positrons trapped in isolated from those trapped in complexes due to very similar CDB ratio curves for both defects, see Fig. 6b.

Figure 6
figure 6

The CDB ratio curves (related to defect-free u-CdTe) (a) experimental data for CdZnTe:Ge in the as-grown state and after annealing in Cd and Te vapour; (b) calculated ratio curves for VCd and VCdGeCd defects. The calculated curves are plotted by dashed lines in the low momentum range (pz < 15 × 10−3m0c) where the accuracy of the momentum distribution calculated by ASTUP is poor.

In spite of extensive research, the basic properties of native point defects in CdTe/CdZnTe are still far from agreed understanding. Both formation and ionization energies are not settled and analyses based on various defect models or experimental data produce incompatible defect characteristics8. Nowadays, it is generally assumed that the difficulties in detecting of native point defects ensue from their reactivity and fast diffusion at elevated temperature. During cooling to room temperature native defects are supposed to migrate and interact with charged impurities. As a result, various types of associates and precipitates are formed58,59. The remaining isolated mostly annihilate during cooling at surface or dislocations and the final room temperature density of comes out very low42. Evidently, such premise is in pure agreement with evaluated defect structure both in CdTe:Cl and CdZnTe:Ge. Analysis of PAS data suggests that all detected are incorporated in various complexes stabilizing them in the lattice.

Detailed characterization of defect structure of CdTe:Cl and CdZnTe:Ge samples summarized in Table 7 raises a query on the unambiguity of the presented defect model. We believe that the comprehensive exploration of CdTe:Cl and CdZnTe:Ge samples, which were studied both in Cd- and in Te-rich state by multiple techniques, LT, CDB and galvanomagnetic measurements, and characterized by GDMS, allowed us the interpreting of measured data by rather unique way. In view of the fact that the defect densities in Te-rich state significantly exceed the densities of obvious residual extrinsic acceptor, we may conclude that is the only candidate for the acceptor defect at these conditions.

Similarly, the intentional doping by Cl or Ge donors supplied the only extrinsic elements with the large enough concentration to arrange the detected density of revealed A-centres and complexes. Nevertheless, suggested interpretation of the defect structure of complexes does not represent the only possibility to create neutral or acceptor-type complex in CdTe and CdZnTe. An analogous defect might be created also without participation of extrinsic atoms taking only native point defects into account. That is the complex of Cd vacancy and Te anti-site defect predicted in ref. 60. Such complex was not identified in experiments yet and its electrical properties are not known. The reason, why we favour the structure of complex related to extrinsic Cl or Ge donors rises from the proximity of the evaluated complex densities in Te-annealed samples and respective element concentrations revealed by GDMS. The density of complexes appears significantly below extrinsic-donor-related complexes and does not manifest in presented measurements. This picture is supported also by theoretical calculations in Table 3 since the complex is characterized positron lifetime which is slightly lower that the lifetime for isolated . But such trend was not observed in experimental lifetimes.

Conclusions

Point defects in CdTe:Cl and CdZnTe:Ge crystals have been characterized by LT and CDB spectroscopies combined with galvanomagnetic measurements. The as-grown CdTe:Cl crystal exhibits larger point defects with open volume comparable with several vacancies. The low temperature LT studies in CdTe:Cl have proved the existence of negatively-charged shallow traps interpreted as Rydberg states associated with Cl-related A-centres. These shallow traps are able to confine positrons at low temperatures only when thermal de-trapping becomes sufficiently small. The lifetime of positrons localized in the shallow traps is shorter than the CdTe bulk lifetime. The as-grown CdZnTe:Ge crystal contains shallow acceptors forming neutral complexes with deep donors. It has been confirmed that Cd-rich annealing of both CdTe:Cl and CdZnTe:Ge crystals suppresses the concentration of , which are created again during subsequent Te-rich annealing.

Additional Information

How to cite this article: Šedivý, L. et al. Positron annihilation spectroscopy of vacancy-related defects in CdTe:Cl and CdZnTe:Ge at different stoichiometry deviations. Sci. Rep. 6, 20641; doi: 10.1038/srep20641 (2016).