Next Article in Journal
Gallate-Based Metal–Organic Frameworks, a New Family of Hybrid Materials and Their Applications: A Review
Next Article in Special Issue
Response of 4H-SiC Detectors to Ionizing Particles
Previous Article in Journal
New Coordination Compounds of CuII with Schiff Base Ligands—Crystal Structure, Thermal, and Spectral Investigations
Previous Article in Special Issue
Depth Profile Analysis of Deep Level Defects in 4H-SiC Introduced by Radiation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Interstitial Carbon–Dioxygen Center in Irradiated Silicon

by
Marianna S. Potsidi
1,
Navaratnarajah Kuganathan
2,3,
Stavros-Richard G. Christopoulos
2,
Alexander Chroneos
2,3,*,
Theoharis Angeletos
1,
Nicholas V. Sarlis
1 and
Charalampos A. Londos
1,*
1
Department of Physics, Solid State Physics Section, National and Kapodistrian University of Athens, Panepistimiopolis Zografos, 15784 Athens, Greece
2
Faculty of Engineering, Environment and Computing, Coventry University, Prior Street, Coventry CV1 5FB, UK
3
Department of Materials, Imperial College London, London SW7 2AZ, UK
*
Authors to whom correspondence should be addressed.
Submission received: 4 October 2020 / Revised: 28 October 2020 / Accepted: 29 October 2020 / Published: 5 November 2020
(This article belongs to the Special Issue Crystalline Materials for Radiation Detection: A New Perspectives)

Abstract

:
We investigated, experimentally as well as theoretically, defect structures in electron irradiated Czochralski-grown silicon (Cz-Si) containing carbon. Infrared spectroscopy (IR) studies observed a band at 1020 cm−1 arisen in the spectra around 300 °C. Its growth occurs concomitantly with the decay out of the well-known vacancy-oxygen (VO) defect, with a Local Vibrational Mode (LVM) at 830 cm−1 and carbon interstitial-oxygen interstitial (CiOi) defect with a LVM at 862 cm−1, in silicon (Si). The main purpose of this work is to establish the origin of the 1020 cm−1 band. One potential candidate is the carbon interstitial-dioxygen (CiO2i) defect since it is expected to form upon annealing out of the CiOi pair. To this end, systematic density functional theory (DFT) calculations were used to predict the lowest energy structure of the (CiO2i) defect in Si. Thereafter, we employed the dipole–dipole interaction method to calculate the vibrational frequencies of the structure. We found that CiO2i defect has an LVM at ~1006 cm−1, a value very close to our experimental one. The analysis and study of the results lead us to tentatively correlate the 1020 cm−1 band with the CiO2i defect.

1. Introduction

The properties and behavior of semiconductors are affected by point defects and impurities introduced in the lattice during crystal growth and material processing. Si mainly contains carbon (C) and oxygen (O), which are unintentionally introduced impurities during crystal growth [1,2]. Upon irradiation, numerous reactions occur, leading to the formation of a variety of carbon-related and oxygen-related defects. Among these, the carbon interstitial (Ci), carbon interstitial-oxygen interstitial (CiOi, carbon interstitial-carbon substitutional (CiCs), and the vacancy-oxygen (VO), which give rise to electrical and optical signals and have been studied both theoretically and experimentally using a range of techniques [1,2,3,4,5,6,7,8,9,10]. Notably, the CiOi, CiCs, and the VO are important recombination centers with adverse effects on the performance of Si-based devices [8,9,10]. Upon thermal annealing, these defects participate in many reactions leading to the production of various complexes. Therefore, it is necessary to know all the properties and behavior of these defects in order to improve the performance of the relative devices.
In the present study, we mainly investigate defects formed upon annealing out of the CiOi pair. The latter defect anneals out mainly by dissociation [1,11]. However, CiOi pairs participate in reactions with other defects where a small percentage of the CiOi migrate as an entity [1]. It has been suggested [12] that in the course of this migration, CiOi reacts with the VO. Both centers anneal out around 300 °C, and when they encounter, they lead to the formation of the carbon substitutional- dioxygen (CsO2i) defect, with a Local Vibrational Mode (LVM) frequency at 1048 cm−1. Alternatively, CsO2i could form according to the reaction [13], Ci + VO2 → CsO2i. Reasonably, upon CiOi annealing, other carbon-related complexes could also form. Such complex is the carbon interstitial-dioxygen (CiO2i) defect [14,15]. In our studies, besides the 1048 cm−1 band, we have seen another one at 1020 cm−1 to arise in the spectra around 300 °C. This band is probably connected with the formation of new second-generation defects formed upon the disappearance of the CiOi pair. A potential candidate for the origin of the 1020 cm−1 band is the CiO2i complex.
In a recent theoretical study [15], the relevant reactions and the migration paths of the interstitial carbon and interstitial oxygen impurities in Si were explored in detail. In particular, a model was presented regarding the formation and dissociation mechanisms of the CiOi and the CiO2i complexes. The suggested [15] reaction paths for the formation of the latter defect involves (i) a further accumulation of oxygen in CiOi (CiOi + Oi → CiO2i) or (ii) a reaction between Ci and O2i (Ci +O2i → CiO2i). Formation mechanism (i) can occur via oxygen migration with an energy barrier of 2.53 eV, or CiOi migration, with a barrier of 2.45 eV. However, the dissociation of CiOi through reaction CiOi → Ci+Oi, has a barrier of 2.29 eV, providing the necessary conditions for formation reaction ii) to occur, ending up to a more stable CiO2i complex, at a barrier of 2.33 eV. Early photoluminescence (PL) studies reported thermal treatments at 350–450 °C, have correlated the decay of the C-line (789.5 meV) of the CiOi defect with the growth of the P-line (767 meV) attributed to the CiO2i defect [1,16]. Recent deep level transient spectroscopy (DLTS) studies have monitored the decay of the CiOi pair and the formation of the CiO2i structure via the observation of the corresponding electronic levels at EV + 0.36 eV and EV + 0.39 eV [17]. These conclusions are in agreement with previous PL and DLTS measurements referring to the investigation of the CiO2i defect [18,19]. It was suggested that the annealing of the CiOi occurs via dissociation into Ci and Oi, where a percentage of the released Ci atoms are trapped by oxygen dimers to form the CiO2i [17,18].
Defects that introduce electronic levels in the gap most possibly affect the functionality of devices by deteriorating in general their performance. It is, therefore, important to identify the origin of any defect signal appearing in the measurements. In this work, we have employed Infrared spectroscopy (IR) measurements coupled with density functional theory (DFT) and dipole–dipole interaction calculations to investigate the 1020 cm−1 band in irradiated Si. The whole analysis of the results has led us to correlate the 1020 cm−1 band with the CiO2i defect.

2. Materials and Methods

2.1. Experimental

We used Czochralski (Cz-Si) p-type (boron-doped) samples with a resistivity of ρ ~ 10 Ω cm. The initial oxygen and carbon concentrations were [Oi] = 1.1 × 1018 and [Cs] = 1.6 × 1017 cm−3, respectively. The samples were irradiated with 2 MeV electrons with a fluence of 1 × 1018 cm−2 at about 80 °C, using the Dynamitron accelerator at JAERI (Takasaki, Japan). After irradiation, the samples were subjected to isochronal anneals at selective temperatures up to ~600 °C in open furnaces. Successive anneals with increased temperature in steps of ΔT ~10 °C and duration Δt = 20 min were conducted. After each annealing step, IR spectra were recorded at room temperature (RT) by means of a Fourier Transform Infrared (FTIR) spectrometer with a resolution of 1 cm−1. The two phonon background absorption was always subtracted from each spectrum by using a Float-zone sample of equal thickness with that of the Cz samples.

2.2. Theoretical

Here, we employed spin-polarized DFT calculations using the Vienna Ab initio Simulation Package (VASP) code [20,21]. Plane wave basis sets and projected augmented wave (PAW) [22] potentials were applied to solve the standard Kohn–Sham (KS) equations. The supercell consisted of 250 Si atomic sites. The generalized gradient approximation (GGA) by Perdew, Burke, and Ernzerhof (PBE) [23] was employed for the exchange-correlation effects. The plane wave basis set extended to a cut-off of 500 eV. The 2 × 2 × 2 Monkhorst-Pack [24] k-point mesh yielded 8 irreducible k-points. Both positions of atoms and lattice parameters were relaxed simultaneously for the Geometry optimizations (constant pressure conditions) using a conjugate gradient algorithm [25]. Convergence criteria were: Force tolerance 0.001 eV/Å and stress tensor 0.002 GPa. The dispersion interactions were accounted for in the form introduced by Grimme et al. [26]. The present calculations were well converged, and the adequacy of the present methodology was discussed in previous work [27,28,29,30,31].

3. Results

3.1. IR Measurements

Figure 1 represents characteristic segments of the IR spectra of electron irradiated Si samples, recorded after irradiation, and at the temperature of 350 °C in the course of the isochronal anneals performed in this experiment. Well-known bands at 830 cm−1 (VO), at 862 cm−1 (CiOi), and a pair at 936, 1020 cm−1 (CiOi(SiI)), were present in the samples immediately after irradiation [30]. After the 350 °C anneal, the first two bands of VO and CiOi diminished substantially, although the pair of bands of the CiOi(SiI) defect disappeared as expected since they were stable up to ~180 °C [32]. Additionally, a set of new bands emerged in the spectra, those at 888 cm−1 (VO2), 1048 cm−1 (CsO2i) [12,13], and another band at 1020 cm−1. The latter band grown in the spectra at ~300 °C in the course of annealing was different from the 1020 cm−1 of the CiOi(SiI) defect appearing in the spectra immediately after irradiation. Its study is the main focus of the present work.
Figure 2 shows the thermal evolution of the VO, VO2 as well as that of the CiOi and 1020 cm−1 bands. The latter band appeared in the spectra in the temperature range of 300 to 550 °C. Its evolution seemed to be connected with the decay of the CiOi defect and a potential candidate for its origin was the CiO2i structure. We have performed DFT calculations in conjunction with dipole–dipole interaction calculations to investigate the possible connection of the 1020 cm−1 band with the CiO2i structure.

3.2. DFT Results

Before proceeding, in order to be an agreement between the notation of DFT outcomes and that used in the dipole–dipole analysis, we will make a few remarks. When used, C refers to the Ci defect, O(1) refers to the Oi atom in the CiOi defect, and O(2) refers to the second Oi atom in the CiO2i defect. Thus, at some points, the CiOi defect is noted C–O(1).
The application of DFT calculations in the semiconductor materials is well established, and they can be used to provide information on the electronic properties and structure of point defects [33]. Considering defect clusters, their exact structure and geometry cannot be easily determined using experiments. This is an area where computational simulation can offer complementary information. To predict the prevalent structure, we considered all possible configurations (nearest neighbor, next nearest neighbor, and further apart) first for the CiOi and thereafter the CiO2i defect. The lowest energy CiOi defect is in excellent agreement with the structure previously reported [34,35,36,37] and references therein. Adding a further Oi leads to the formation of the CiO2i defect (refer to Figure 3). It should be stressed that the reported structure is the lowest energy one after systematically examining all possible configurations of the CiO2i defect in the Si lattice. This structure was thereafter adopted as the most possible one for our calculations.
For completeness, we also considered the density of states (DOS) plots and charge density plots (refer to Figure 4). There is a small gap state noted just below the Fermi level (refer to Figure 4a). This state belongs to the carbon interstitial, as confirmed by the atomic DOS of carbon (refer to Figure 4b). States belonging to the oxygen interstitial are mainly located in the deeper level (between 0 eV and 1 eV) (refer to Figure 4c,d). Charge density plots of oxygen di-interstitials are shown in Figure 4g,h. Decomposed charge density plot associated with carbon interstitial is delocalized over Si atoms it bonded to. Decomposed charge density plots of oxygen interstitials delocalized over the square planar ring.

3.3. LVM Estimations via the Dipole–Dipole Interaction Method

Figure 5 is a simplified structure model configuration of the CiO2i defect derived by DFT calculations. In essence, Figure 5 displays part of the lattice of Figure 3 containing the C and O(1) atoms settled in the (110) plane and the second oxygen interstitial atom, O(2). The C–O(1) direction forms an approximately 20° deeping angle with the [110] direction (the Si4–Si5 direction). Both C–O(1) and O(2) are IR active with reported [5] apparent charges and can be considered as two oscillating interacting dipoles.
Regarding the C–O(1) dipole, C and O(1) atoms are taken as one particle with mass corresponding to the reduced mass of oxygen and carbon atom μ, located at the center of mass, at distance d from O(1) atom. The reduced mass μ is derived from the relation
μ = m O m C m O + m C
while the distance d from the relation
d   = m C m O + m C d C O ( 1 )
where   m O = 16   a m u and m C = 12   a m u are the masses of the oxygen and carbon atom, respectively and d C O ( 1 ) is the distance between the C and O(1) atoms. The calculated values for μ and d, are μ = 11.38 × 10−27 kg and d = 1.07 Å, respectively.
The effective charge of C–O(1) is concentrated on the C–O(1) reduced mass and is equal to [5] Z C O ( 1 ) = 2.4   | e | , where | e | is the electron charge. The effective charge concentrated on the O(2) atom is equal to [5] Z O ( 2 ) = 4.1   | e | , where | e | is the electron charge.
To estimate the LVM frequencies of the CiO2i defect, we applied a previously used procedure [38,39] based on the interaction of the C–O(1) and O(2) dipoles. The force constant of the perturbed oscillating C–O(1) entity is K C O ( 1 ) and is given by the relation
K C O ( 1 ) = μ · ( ω C O ( 1 ) ) 2
where ω C O ( 1 ) are the LVM frequencies of the C–O(1) defect. Similarly, the force constant of the oscillating O(2) atom, K O ( 2 ) , is given by the relation
K O ( 2 ) =   m O · ( ω O ( 2 ) ) 2
and corresponds to the ω O ( 2 ) =1107 cm−1 LVM frequency of the oxygen interstitial atom [2]. In this analysis, we use the values of K C O ( 1 ) corresponding to the strongest experimentally detected LVM line at 862 cm−1 and the reported line [1,40] at 1116 cm−1, both attributed to carbon-related modes.
We adopt the C–O(1) direction for the oscillation of the C–O(1) dipole and the [ 1 1 ¯ 1 ¯ ] for the oscillation [41] of the O(2) dipole. With the help of Figure 5, the expressions of the dipole moments p C O ( 1 ) and p O ( 2 ) become
p C O ( 1 ) = Z C O ( 1 ) q 1 q ^ 1 p O ( 2 ) = Z O ( 2 ) q 2 q ^ 2
where q ^ 1 ,   q ^ 2 are the unit vectors along the C–O(1) and [ 1 1 ¯ 1 ¯ ] directions, respectively. The potential energy of the interacting dipole moments is given by [42]
U int = p C O ( 1 ) ·   p O ( 2 ) 3 ( n ^ ·   p C O ( 1 ) ) ( n ^ ·   p O ( 2 ) ) d μ O ( 2 ) 3
where n ^ is the unit vector along the direction that connects the two dipoles and d μ O ( 2 )   = 3.36   Å is the distance between them. The motion of the two dipoles is described by the effective Hamiltonian
H = 1 2 μ q ˙ 1 2 + 1 2   m O q ˙ 2 2 + 1 2 K C O ( 1 ) q 1 2 + 1 2 K O ( 2 ) q 2 2 + λ q 1 q 2
By comparing Equations (6) and (7), we obtain the q-independent part λ, to be equal to λ = 138 J/m2. The Hamiltonian of Equation (7) has two normal modes with frequencies
ω C i O 2 i = 1 2 [ ω C O ( 1 ) 2 + ω O ( 2 ) 2 ± ( ω C O ( 1 ) ω O ( 2 ) ) 2 + 4   λ 2 μ ·   m O ]
where “+” preceding the inner square root corresponds to the antisymmentric normal mode and “−“ to the symmetric one. For ω C O ( 1 ) =   862 cm−1 and ω O ( 2 ) = 1107 cm−1 we find ω 1 ,   C i O 2 i ( a n t ) =1098 cm−1 and ω 1 ,   C i O 2 i ( s y m ) =872 cm−1. For ω C O ( 1 ) =   1116 cm−1 and ω O ( 2 ) = 1107 cm−1 we find ω 2 ,   C i O 2 i ( a n t ) =1205 cm−1 and ω 2 ,   C i O 2 i ( s y m ) =1006 cm−1. Among them, we consider the band at 1006 cm−1 as the most possible candidate to be related to our experimental findings, as discussed below.

4. Discussion

The calculated frequency at 1006 cm−1 standing for the symmetric mode of the CiO2i related to the 1116 cm−1 carbon-related mode of the C–O(1) defect lies very close to the experimental line at 1020 cm−1, supporting the attribution of this line to the CiO2i defect. Since the calculated dot product of the dipole moment vectors is positive, one would expect the symmetric modes at 872 and 1006 cm1 to give stronger IR signals than the antisymmetric ones, at 1098 and 1205 cm1. In our spectra, only the symmetric mode at 1006 cm1 is detectable. The other band at 872 cm1 is not detected in our experiment. This is possibly because its weak signal is masked by the stronger signal of the VO2 defect (Figure 1) in the same region. It is worth noting that if both dipoles oscillated along the direction connecting them, the theoretical outcome through the dipole–dipole interaction process would have provided a better fit (1021 cm−1) to the experimental value at 1020 cm−1.
The present calculations provide support for the notion that the 1020 cm−1 line can originate from the CiO2i defect. Regarding this correlation, it deserves noting that early IR studies of the evolution of carbon and its loss from solution, upon thermal treatments of Cz-Si at 450 °C, have revealed the formation of a band at 1026 cm−1 attributed to a carbon-oxygen dimer complex [43,44,45].
An important point: In our studies of electron irradiated silicon, the 1020 cm−1 band is formed around 300 °C upon the decay of the 862 cm−1 band of the CiOi pair. However, in proton-irradiated thermally treated Si, the formation of the CiO2i occurs at temperatures above 400 °C according to the reaction Ci +O2i → CiO2i [17,18,19]. The suggested mechanism in the latter reports involves the dissociation of the CiOi and the capture of the liberated carbon interstitials by oxygen dimers. Profoundly, at these temperatures the availability of oxygen dimers plays an important role in the validity of the above reaction. In our case of electron irradiated silicon, CiO2i emerges in the spectra around 300 °C. However, the concentration of dimers at ~300 °C is practically negligible [2], and the reaction Ci +O2i → CiO2i could not account for the formation of the CiO2i. Other reaction channels should govern the formation of the CiO2i in our case. One suggested reaction is the following: It is known that CiOi anneals out mainly by dissociation (CiOi → Ci + Oi), but a small percentage of them could migrate in the lattice as an entity [1,15], being trapped by Oi atoms. This process leads to the formation of CiO2i structure (Oi + CiOi → CiO2i). Additionally, another reaction mechanism could be envisaged. It is known that among the reactions that VO pair participates upon annealing, the main two are the following [7]: VO + Oi → VO2 and VO + SiI → Oi. In the latter reaction, sources of the participating self-interstitials are self-interstitial clusters formed in the course of the heavy irradiation of the Cz-Si [2]. In addition, Oi could be additionally provided [7] by the dissociation of VO. Since VO anneals out in parallel with CiOi, the following scheme is suggested. The liberated Oi atoms, in the course of VO dissociation or/and its destruction by self-interstitials, trap CiOi pairs leading to CiO2i complexes (VO + SiI → Oi, VO → V + Oi, and then Oi + CiOi → CiO2i). Apparently, the above two suggested reaction channels for the formation of the CiO2i defect may occur in parallel. Such reactions could account for the formation of the CiO2i in irradiated Si and explain its appearance at lower temperatures than those in thermally treated Si.
Regarding the differences in the formation temperature of the CiO2i defect between our electron irradiated samples and proton irradiated samples [17,18,19], the following should be noted: Proton irradiation generates higher order of damage than electron irradiation. This might result in more stable defects, which may need a higher temperature to anneal out. Additionally, the detected signal in our IR studies could refer to different states of the same defect reported in previous photoluminescence (PL) and DLTS measurements [17,18,19]. Further clarification of this issue requires future work.
Radiation defects in semiconductors attract a lot of interest, especially for technological purposes. Their presence in the Si lattice, in the course of material processing for special applications as for instance, radiation detectors, affect the device operation. Apparently, any information for such defects may improve the device functionality since it provides opportunities for their control. The present study is an attempt to determine the origin of an IR signal at 1020 cm−1. We have connected it with a second-order generation carbon-oxygen related defect, that is the CiO2i complex. In future work we shall use higher irradiation fluencies to study the annealing of the CiO2i defect and the formation of larger complexes in this family of defects as the CiO3i structure, recently predicted theoretically [15].

5. Conclusions

The annealing of defects in irradiated Cz-Si revealed a new IR band at 1020 cm−1 emerging in the spectra around 300 °C upon the decay of the 862 cm−1 band of the CiOi defect. PL and DLTS data have previously correlated certain signals with the formation of a CiO2i defect grown in the spectra upon the decay of the CiOi defect. In the present study, we employed systematic DFT calculations to gain an understanding of the lowest energy structure of the CiO2i defect. Thereafter, the dipole–dipole interaction calculations have estimated an LVM frequency at 1006 cm−1 related to this center. This value lies very close to the experimental one at 1020 cm−1. We have assigned the latter band to the CiO2i defect.

Author Contributions

Formal analysis, T.A., N.V.S. and C.A.L.; methodology, M.S.P., N.K., S.-R.G.C., and C.A.L.; supervision, A.C.; writing—original draft, M.S.P., N.K., S.-R.G.C., and C.A.L.; writing—review and editing, A.C. and N.V.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Acknowledgments

T. Angeletos is grateful to the A. S. Onassis Foundation for financial support through his Ph.D. scholarship (Grant No. G ZL 001-1/2015-2016).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Davies, G.; Newman, R.C. Carbon in Monocrystalline Silicon. In Handbook on Semiconductors; Completely Revised ed.; Moss, T.S., Mahajan, S., Eds.; Elsevier Science, B.V.: Amsterdam, The Netherlands, 1994; Volume 3, pp. 1557–1635. [Google Scholar]
  2. Newman, R.C.; Jones, R. Diffusion of Oxygen in Silicon. In Semiconductors and Semimetals: Oxygen in Silicon; Shimura, F., Ed.; Academic Press: Orlando, FL, USA, 1994; Volume 42, pp. 289–352. [Google Scholar]
  3. Londos, C.A. Annealing Studies of Defects Pertinent to Radiation Damage in Silicon. Phys. Stat. Sol. 1987, 102, 639–644. [Google Scholar] [CrossRef]
  4. Ferenczi, G.; Londos, C.A.; Pavelka, T.; Somogyi, M. Correlation of the concentration of the carbon-associated radiation damage levels with the total carbon concentration in silicon. J. Appl. Phys. 1988, 63, 183–189. [Google Scholar] [CrossRef]
  5. Chappell, S.P.; Claybourn, M.; Newman, R.C.; Barraclough, K.G. The production of Ci-Oi complexes during irradiation of Czochralski silicon at 130 K. Semicond. Sci. Technol. 1988, 3, 1047–1051. [Google Scholar] [CrossRef]
  6. Watkins, G.D.; Corbett, J.W. Defects in Irradiated Silicon. I. Electron Spin Resonance of the Si-A Center. Phys. Rev. 1961, 121, 1001–1014. [Google Scholar] [CrossRef]
  7. Londos, C.A.; Andrianakis, A.; Sgourou, E.N.; Emtsev, V.V.; Ohyama, H. IR studies of the impact of Ge doping on the successive conversion of VOn defects in Czochralski-Si containing carbon. J. Appl. Phys. 2011, 109, 1–8. [Google Scholar] [CrossRef] [Green Version]
  8. Tsuchiya, D.; Sueoka, K.; Yamamoto, H. Density Functional Theory Study on Defect Behavior Related to the Bulk Lifetime of Silicon Crystals for Power device Application. Phys. Status Solidi A 2019, 216, 1–17. [Google Scholar] [CrossRef]
  9. Brotherton, S.D.; Bradley, P.J. Defect production and lifetime control in electron and γ-irradiated silicon. Appl. Phys. 1982, 53, 5720–5732. [Google Scholar] [CrossRef]
  10. Siemieniec, M.; Niedernostheide, F.-J.; Schlze, H.-J.; Sudkamp, W.; Kellner-Werdehausen, U.; Lutz, J. Irradiation-Induced Deep Levels in Silicon for Power Device Tailoring. J. Electroch. Soc. 2006, 153, G108–G118. [Google Scholar] [CrossRef]
  11. Svensson, B.G.; Lindström, J.L. Annealing Studies of the 862 cm−1 Infrared Band in Silicon. Phys. Stat. Sol. 1987, 95, 537–542. [Google Scholar] [CrossRef]
  12. Inoue, N.; Ohyama, H.; Goto, Y.; Sugiyama, T. Quantitative analysis of complexes in electron irradiated CZ silicon. Physica B 2007, 401, 477–482. [Google Scholar] [CrossRef]
  13. Murin, L.I.; Markevich, V.P.; Lindstrom, J.L.; Kleverman, M.; Hermansson, J.; Hallberg, T.; Svensson, B.G. Carbon-Oxygen-Related Complexes in Irradiated and Heat-Treated Silicon: IR Absorption Studies. Solid State Phenom. 2002, 82, 57–62. [Google Scholar] [CrossRef]
  14. Ewels, C.P.; Jones, R.; Oberg, S. Oxygen-Carbon, Oxygen-Nitrogen and Oxygen Dimer Defects in Silicon. In Early Stages of Oxygen Precipitation in Silicon; Jones, R., Ed.; Kluwer Academic Publishers: London, UK, 1996; Volume 17, pp. 141–162. [Google Scholar]
  15. Ayed, H.M.; Monakhov, E.V.; Coutinho, J. Formation and dissociation reactions of complexes involving interstitial carbon and oxygen defects in silicon. Phys. Rev. Mater. 2020, 4, 064601. [Google Scholar] [CrossRef]
  16. Kürner, W.; Sauer, R.; Dornen, A.; Thonke, K. Structure of the 0.767-eV oxygen-carbon luminescence defect in 450 oC thermally annealed Czochralski-grown silicon. Phys. Rev. B 1989, 39, 13327–13337. [Google Scholar] [CrossRef] [PubMed]
  17. Ayedh, H.M.; Grigorev, A.A.; Galeckas, A.; Svensson, B.G.; Monakhov, E.V. Annealing Kinetics of the Interstitial Carbon-Dioxygen Complex in Silicon. Phys. Status Solidi A 2019, 216, 1–5. [Google Scholar] [CrossRef]
  18. Ganagona, N.; Raeissi, B.; Vines, L.; Monakhov, E.V.; Svensson, B.G. Defects in p-type Cz-silicon irradiated at elevated temperatures. Phys. Status Solidi C 2012, 9, 2009–2012. [Google Scholar] [CrossRef]
  19. Raeissi, B.; Ganagona, N.; Galeckas, A.; Monakhov, E.V.; Svensson, B.G. PL and DLTS analysis of carbon-related centers in irradiated p-type Cz-Si. Solid State Phenom. 2014, 205, 224–227. [Google Scholar] [CrossRef]
  20. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186. [Google Scholar] [CrossRef]
  21. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775. [Google Scholar] [CrossRef]
  22. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef] [Green Version]
  23. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef] [Green Version]
  24. Monkhorst, H.J.; Pack, J.D. Special points for Brillouin-zone integration. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  25. Press, W.H.; Teukolsky, S.A.; Vetterling, W.T.; Flannery, B.P. Numerical Recipes in C: The Art of Scientific Computing, 2nd ed.; Cambridge University Press: Cambridge, UK, 1992. [Google Scholar]
  26. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 1–19. [Google Scholar] [CrossRef] [Green Version]
  27. Chroneos, A.; Bracht, H.; Grimes, R.W.; Uberuaga, B.P. Phosphorus clustering in germanium-rich silicon germanium. Mater. Sci. Eng. B 2008, 154, 72–75. [Google Scholar] [CrossRef]
  28. Chroneos, A.; Jiang, C.; Grimes, R.W.; Schwingenschlögl, U.; Bracht, H. E centers in ternany Si1−x−yGexSny random alloys. Appl. Phys. Lett. 2009, 95, 1–3. [Google Scholar] [CrossRef] [Green Version]
  29. Chroneos, A.; Londos, C.A. Interaction of A-centers with isovalent impurities in silicon. J. Appl. Phys. 2010, 107, 1–4. [Google Scholar] [CrossRef] [Green Version]
  30. Sgourou, E.N.; Timerkaeva, D.; Londos, C.A.; Aliprantis, D.; Chroneos, A.; Caliste, D.; Pochet, P. Impact of isovalent doping on the trapping of vacancy and interstitial defects in silicon. J. Appl. Phys. 2013, 113, 113506. [Google Scholar] [CrossRef] [Green Version]
  31. Wang, H.; Chroneos, A.; Londos, C.A.; Sgourou, E.N.; Schwingenschlögl, U. A-centers in silicon with hybrid density functional theory. Appl. Phys. Lett. 2013, 103, 052101. [Google Scholar] [CrossRef] [Green Version]
  32. Londos, C.A.; Sgourou, E.N.; Chroneos, A.; Emtsev, V.V. Carbon, oxygen and intrinsic defect interactions in germanium-doped silicon. Semicond. Sci. Technol. 2011, 26, 1–6. [Google Scholar] [CrossRef]
  33. Chroneos, A.; Sgourou, E.N.; Londos, C.A.; Schwingenschlögl, U. Oxygen defect processes in silicon and silicon germanium. Appl. Phys. Rev. 2015, 2, 021306. [Google Scholar] [CrossRef] [Green Version]
  34. Khirunenko, L.I.; Sosnin, M.G.; Pomozov, Y.V.; Murin, L.I.; Markevich, V.P.; Peaker, A.R.; Almeida, L.M.; Coutinho, J.; Torres, V.J.B. Formation of interstitial carbon-interstitial oxygen complexes in silicon: Local vibrational mode spectroscopy and density functional theory. Phys. Rev. B 2008, 78, 155203. [Google Scholar] [CrossRef]
  35. Coutinho, J.; Jones, R.; Briddon, P.R.; Ӧberg, S.; Murin, L.I.; Markevich, V.P.; Lindstrӧm, J.L. Interstitial carbon-oxygen center and hydrogen related shallow thermal donors in Si. Phys. Rev. B 2001, 65, 014109. [Google Scholar] [CrossRef]
  36. Backlund, D.J.; Estreicher, S.K. C4 defect and its precursors in Si: First principles theory. Phys. Rev B 2008, 77, 1–8. [Google Scholar] [CrossRef]
  37. Wang, H.; Chroneos, A.; Londos, C.A.; Sgourou, E.N.; Schwingenschlögl, U. Carbon related defects in irradiated silicon revisited. Sci. Rep. 2014, 4, 1–9. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Sarlis, N.V.; Londos, C.A.; Fytros, L.G. Origin of infrared bands in neutron-irradiated silicon. J. Appl. Phys. 1997, 81, 1645–1650. [Google Scholar] [CrossRef] [Green Version]
  39. Potsidi, M.S.; Londos, C.A. The CiCs(SiI) defect in silicon: An infrared spectroscopy study. J. Appl. Phys. 2006, 100, 1–4. [Google Scholar] [CrossRef]
  40. Davies, G.; Hayama, S.; Hao, S.; Bech Nielsen, B.; Coutinho, J.; Sanati, M.; Estreicher, S.K.; Itoh, K.M. Host isotope effects on midinfrared optical transitions in silicon. Phys. Rev. B 2005, 71, 1–7. [Google Scholar] [CrossRef] [Green Version]
  41. Mc Cluskey, M.D. Local vibrational modes of impurities in semiconductors. J. Appl. Phys. 2000, 87, 3593–3617. [Google Scholar] [CrossRef]
  42. Jackson, J. Classical Electrodynamics, 2nd ed.; Wiley: New York, NY, USA, 1975; p. 136. [Google Scholar]
  43. Bean, A.R.; Newman, R.C. The effect of carbon on thermal donor formation in heat treated pulled silicon crystals. J. Phys. Chem. Solids 1972, 33, 251–268. [Google Scholar] [CrossRef]
  44. Newman, R.C.; Oates, A.S.; Livingston, F.M. Self-interstitials and thermal donor formation in silicon: New measurements and a model for the defects. J. Phys. C Solid State Phys. 1983, 16, L667–L674. [Google Scholar] [CrossRef]
  45. Lindström, J.L.; Werman, H.; Oehrlein, G.S. Thermal donors and carbon-Oxygen defects in silicon. Phys. Status Solid A 1987, 99, 581–591. [Google Scholar] [CrossRef]
Figure 1. Segments of IR spectra of electron irradiated Si samples immediately after irradiation and at the characteristic T of 350 °C during the isochronal anneal sequence.
Figure 1. Segments of IR spectra of electron irradiated Si samples immediately after irradiation and at the characteristic T of 350 °C during the isochronal anneal sequence.
Crystals 10 01005 g001
Figure 2. The thermal evolution of the 830 cm−1 (VO), 862 cm−1 (CiOi), 888 cm−1 (VO2), and the 1020 cm−1 band.
Figure 2. The thermal evolution of the 830 cm−1 (VO), 862 cm−1 (CiOi), 888 cm−1 (VO2), and the 1020 cm−1 band.
Crystals 10 01005 g002
Figure 3. Schematic representation of the energetically favorable CiO2i defect. Bond distance values are: d1 = 1.665 Å, d2 = 1.626 Å, d3 = 1.711 Å, d4 = 1.818 Å, d5 = 1.765 Å, d6 = 1.781 Å, d7 = 1.782 Å, d8 = 1.809 Å, d9 = 2.571 Å. Angle values are: Si1–O(2)–Si2 = 140.85°, O(2)–Si2–O(1) = 99.04°, Si2–O(1)–Si3 = 140.24°, Si2–O(1)–Si5 = 129.45°, O(1)–Si3–C = 88.51°, O(1)–Si5–C = 88.25°, O(1)–Si3–Si5 = 44.72°, O(1)–Si5–Si3 = 44.98°, Si5–Si3–C = 43.79°, Si3–Si5–C = 43.27°, Si3–C–Si4 = 125.69°, Si5–C–Si4 = 141.37°.
Figure 3. Schematic representation of the energetically favorable CiO2i defect. Bond distance values are: d1 = 1.665 Å, d2 = 1.626 Å, d3 = 1.711 Å, d4 = 1.818 Å, d5 = 1.765 Å, d6 = 1.781 Å, d7 = 1.782 Å, d8 = 1.809 Å, d9 = 2.571 Å. Angle values are: Si1–O(2)–Si2 = 140.85°, O(2)–Si2–O(1) = 99.04°, Si2–O(1)–Si3 = 140.24°, Si2–O(1)–Si5 = 129.45°, O(1)–Si3–C = 88.51°, O(1)–Si5–C = 88.25°, O(1)–Si3–Si5 = 44.72°, O(1)–Si5–Si3 = 44.98°, Si5–Si3–C = 43.79°, Si3–Si5–C = 43.27°, Si3–C–Si4 = 125.69°, Si5–C–Si4 = 141.37°.
Crystals 10 01005 g003
Figure 4. (a) Total DOS plot, (b) atomic DOS plot of carbon interstitial (C), (c) atomic DOS plot of the first oxygen interstitial (O(1)), (d) atomic DOS plot of the second oxygen interstitial (O(2)), (e) total charge density plot; decomposed charge density plots associated with (f) C, (g) O(1), and (h) O(2). Black circle refers to a carbon atom, and red circles refer to oxygen atoms.
Figure 4. (a) Total DOS plot, (b) atomic DOS plot of carbon interstitial (C), (c) atomic DOS plot of the first oxygen interstitial (O(1)), (d) atomic DOS plot of the second oxygen interstitial (O(2)), (e) total charge density plot; decomposed charge density plots associated with (f) C, (g) O(1), and (h) O(2). Black circle refers to a carbon atom, and red circles refer to oxygen atoms.
Crystals 10 01005 g004
Figure 5. The CiO2i configuration. The blue hexagon corresponds to the reduced mass μ, of the C–O(1) dipole. Blue arrows indicate the oscillation directions of the C–O(1) and the O(2) dipoles.
Figure 5. The CiO2i configuration. The blue hexagon corresponds to the reduced mass μ, of the C–O(1) dipole. Blue arrows indicate the oscillation directions of the C–O(1) and the O(2) dipoles.
Crystals 10 01005 g005
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Potsidi, M.S.; Kuganathan, N.; Christopoulos, S.-R.G.; Chroneos, A.; Angeletos, T.; Sarlis, N.V.; Londos, C.A. The Interstitial Carbon–Dioxygen Center in Irradiated Silicon. Crystals 2020, 10, 1005. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst10111005

AMA Style

Potsidi MS, Kuganathan N, Christopoulos S-RG, Chroneos A, Angeletos T, Sarlis NV, Londos CA. The Interstitial Carbon–Dioxygen Center in Irradiated Silicon. Crystals. 2020; 10(11):1005. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst10111005

Chicago/Turabian Style

Potsidi, Marianna S., Navaratnarajah Kuganathan, Stavros-Richard G. Christopoulos, Alexander Chroneos, Theoharis Angeletos, Nicholas V. Sarlis, and Charalampos A. Londos. 2020. "The Interstitial Carbon–Dioxygen Center in Irradiated Silicon" Crystals 10, no. 11: 1005. https://0-doi-org.brum.beds.ac.uk/10.3390/cryst10111005

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop