Expand this Topic clickable element to expand a topic
Skip to content
Optica Publishing Group

Dual role of Sb ions as electron traps and hole traps in photorefractive Sn2P2S6 crystals

Open Access Open Access

Abstract

Doping photorefractive single crystals of Sn2P2S6 with antimony introduces both electron and hole traps. In as-grown crystals, Sb3+ (5s2) ions replace Sn2+ ions. These Sb3+ ions are either isolated (with no nearby perturbing defects) or they have a charge-compensating Sn2+ vacancy at a nearest-neighbor Sn site. When illuminated with 633 nm laser light, isolated Sb3+ ions trap electrons and become Sb2+ (5s25p1) ions. In contrast, Sb3+ ions with an adjacent Sn vacancy trap holes during illumination. The hole is primarily localized on the (P2S6)4− anionic unit next to the Sb3+ ion and Sn2+ vacancy. These trapped electrons and holes are thermally stable below ∼200 K, and they are observed with electron paramagnetic resonance (EPR) at temperatures below 150 K. Resolved hyperfine interactions with 31P, 121Sb, and 123Sb nuclei are used to establish the defect models.

© 2016 Optical Society of America

1. Introduction

Single crystals of Sn2P2S6 (often referred to as SPS) have interesting and useful photorefractive properties [1–4]. Doping with antimony (Sb) ions has been shown to significantly improve the photorefractive response of the SPS crystals, especially in the 530-675 nm region [5–11]. Of special interest is the increase in the photorefractive sensitivity of Sb-doped SPS when there is a pre-exposure to light with wavelengths near the optical band edge [5,9]. Recent observations of the transmittance dynamics (i.e., temporal dependence) of a red testing beam after terminating an intense green “pre-illumination” beam suggest that at least two distinct primary traps with different thermal stabilities are present in the Sb-doped crystals [5]. To better understand these photoinduced results, the identification and characterization of all Sb-related optically active defects in SPS is needed.

In an SPS crystal, Sb3+ (5s2) ions easily substitute for Sn2+ (5s2) ions. These ions have the same electronic structure and similar ionic size (for sixfold coordination, the ionic radius of the Sb3+ and Sn2+ ions are 0.76 and 1.18 Å, respectively). In an otherwise stoichiometric as-grown crystal with no additional dopants present, Sn2+ vacancies are expected to provide charge compensation for the Sb3+ ions. Specifically, there will be one doubly ionized Sn2+ vacancy for every two Sb3+ ions, thus ensuring the overall neutrality of the crystal. These Sb3+ ions and Sn2+ vacancies may be well separated and thus act as isolated electron and hole traps, or they may thermally migrate to each other and form a close-associate pair (i.e., a donor-acceptor pair) that acts only as a hole trap. A recent electron paramagnetic resonance (EPR) study has shown that isolated Sb3+ ions trap an electron when an SPS crystal is illuminated at low temperature with light near the band edge [12]. The resulting Sb2+ (5s25p1) ions are stable for sample temperatures below approximately 200 K. Their thermal decay at higher temperatures is described by second-order kinetics with an activation energy of 420 meV [13]. Also, EPR has shown that isolated Sn2+ vacancies trap a hole during illumination at low temperatures [14]. Other defects in SPS identified by EPR include sulfur vacancies [15] and Sn3+ holelike small polarons [16].

We use EPR in the present paper to experimentally verify that the defect complex consisting of an Sb3+ ion and an adjacent doubly ionized Sn2+ vacancy serves as a hole trap in SPS crystals. The Sb3+ ion replaces a Sn2+ ion and has an effective charge of + 1 (relative to the regular lattice), whereas a doubly ionized Sn vacancy has an effective −2 charge. Before trapping the hole, this complex is a (Sb-VSn) defect with an effective −1 charge. After the hole is trapped, the complex becomes a neutral (Sb-VSn)0 defect. The trapped hole is localized primarily on the one (P2S6)4− unit that is adjacent to both the Sb3+ ion and the Sn vacancy. This anionic unit has a large negative charge and strongly attracts the hole. Resolved hyperfine interactions with 31P, 121Sb, and 123Sb nuclei in the EPR spectra provide the key information that establishes the ground-state model of this Sb-related trapped hole center in SPS crystals.

2. Experimental details

The Sn2P2S6 crystals have a monoclinic structure (space group Pn and point group m) below the paraelectric-ferroelectric phase change at 64 °C. At room temperature, the lattice constants are a = 9.378 Å, b = 7.488 Å, c = 6.513 Å, and β = 91.15°, with the b axis perpendicular to the mirror plane of the crystal [17]. The (P2S6)4− anionic groups exhibit strong covalent behavior and are separated by the more ionic Sn2+ ions [18].

A large single crystal of SPS with dimensions on the order of centimeters was grown by the vertical Bridgman method at Uzhgorod National University (Ukraine). A sample for our EPR experiments (2.0 × 2.5 × 2.5 mm3) was then cut from this larger boule. Reference [14] describes details of the growth process. Starting materials (Sn1.93P2.03S6.04) for this particular growth run were chosen to produce a Sn deficient crystal. Thus, a large concentration of Sn vacancies is present. Although not intentionally doped, this crystal also has a significant concentration of Sb impurities, most likely coming from the starting materials. In addition to Sn vacancies and Sb3+ impurities, sulfur vacancies are present in this as-grown Sn-deficient crystal.

A Bruker EMX spectrometer operating near 9.392 GHz was used to obtain the EPR spectra. The sample temperature was controlled in the 40-220 K region with an Oxford gas flow system. There were no EPR signals in the as-grown crystal at room temperature or at low temperature (when cooled in the dark), as the donors and acceptors are all in nonparamagnetic charge states with no unpaired spins. These defects are converted to paramagnetic charge states during exposure to 633 nm light from a He-Ne laser or 510 nm light from a Xe lamp while the sample is held at low temperature in the microwave cavity (the optical band edge of SPS is near 530 nm at room temperature and shifts to near 500 nm at lower temperature). The SPS crystal was glued to the polished bottom face of a 4 mm diameter and 24 cm long quartz rod inserted into the Oxford cryostat and microwave cavity (the sample and the lower 4 cm of the rod were in direct contact with flowing cold nitrogen or helium gas while the top of the rod remained at room temperature). During an optical illumination, light entering the polished top face of the rod (i.e., light pipe) was transmitted to the cold sample. The use of the quartz rod as a light guide led to uniform filling of the bottom face of the rod with light and to the loss of any polarization of the incident light. For illumination at 510 nm, the output of a 150 W ozone-free Xe arc lamp (Oriel Model 6255) was passed through a 510 nm band-pass interference filter with 10 nm bandwidth (Thorlabs FB510-10). Incident 510 nm light power at the top face of the quartz rod was 5 mW. For illumination at 633 nm, a cw He-Ne laser (Spectra-Physics Model 106-1) operating at a power level of 12 mW was used. Duration of exposures to the 510 or 633 nm light was typically 5 min (thus ensuring that the effects of the illumination on the defects had saturated).

3. Results

3.1 EPR spectra

Figure 1 shows the EPR spectrum obtained from an SPS crystal containing large concentrations of Sb ions and Sn vacancies. While being held at 42 K, the sample was initially exposed to 633 nm laser light. The light was then removed and the sample was warmed to 125 K and held at this temperature for 2 min to reduce interfering signals from isolated Sn and S vacancies [14,15]. These ramps up and down in temperature each took approximately 5 min. After returning to 42 K following this brief warming, the spectrum in Fig. 1 was taken with no further illumination.

 figure: Fig. 1

Fig. 1 EPR spectrum from an SPS crystal taken at 42 K with the magnetic field along the a axis. Blue stick diagrams identify lines from the Sb2+ ions (trapped electrons) and red stick diagrams identify lines from (Sb-VSn)0 complexes (trapped holes). Hyperfine lines from the 121Sb and 123Sb nuclei are present for both defects, as well as 31P hyperfine triplets for the (Sb-VSn)0 complexes.

Download Full Size | PDF

Two S = 1/2 Sb-related defects, with widely split hyperfine patterns, are identified in Fig. 1 by “stick” diagrams. Additional lines in the center region, between 322 and 345 mT, are from the remaining Sn and S vacancies. Isolated Sb2+ ions (trapped electrons) are indicated by the blue stick diagrams and (Sb-VSn)0 complexes (trapped holes) are indicated by the red stick diagrams. The large number of lines in each set are caused by the hyperfine interactions of the unpaired electron with the magnetic moments of the 121Sb and 123Sb nuclei (121Sb nuclei are 57.3% abundant and their I = 5/2 nuclear spin gives six hyperfine lines whereas 123Sb nuclei are 42.7% abundant and their I = 7/2 nuclear spin gives eight hyperfine lines). Nuclear magnetic moments for the Sb isotopes are µ(121Sb) = + 3.3634βn and µ(123Sb) = + 2.5498βn (where βn is the nuclear magneton). The 121Sb and 123Sb hyperfine parameters for a specific defect are proportional to their respective µ/I values. The isolated Sb2+ set of EPR lines in Fig. 1 (blue stick diagrams) has been previously reported [12]. In the present paper, our focus is on the newly observed (Sb-VSn)0 trapped-hole defects.

In Fig. 1, each Sb hyperfine line in the (Sb-VSn)0 spectrum (the red stick diagrams) is a set of three closely spaced lines with approximate intensity ratios of 1:2:1. These triplets are caused by nearly equal hyperfine interactions with two 31P nuclei (the nuclear spin of 31P is I = 1/2 and they are 100% abundant). The triplet sets of lines strongly suggest that the wave function for the unpaired spin has significant overlap with one (P2S6)4− anionic unit (and thus with its two 31P nuclei), as well as the interaction with one Sb ion at a Sn site. The recent study of the EPR spectrum from singly ionized Sn vacancies in SPS shows similar hyperfine interactions with two 31P nuclei [14]. Table 1 provides a comparison of the 31P hyperfine interactions for the two trapped hole centers when the magnetic field is along the three crystallographic directions. For the (Sb-VSn)0 center, any difference in the strength of the interactions with the two 31P nuclei is not resolved (i.e., the difference in these interactions is less than the linewidth). In the case of the isolated Sn vacancy, the two 31P interactions are not equal in magnitude and vary by at most a factor of three [14]. Both defects have a Sn vacancy as the primary element in their models, thus it is not surprising that the trapped hole in each case has similar interactions with the two 31P nuclei in the one adjacent (P2S6)4− unit and that the magnitudes of these 31P interactions are comparable for the two defects. We take the values of the 31P hyperfine interactions in Table 1 to be compelling evidence that the new (Sb-VSn)0 center consists of both a Sn vacancy and a localized hole, as well as the Sb ion. For perspective, hyperfine interactions with the two 31P nuclei in the singly ionized sulfur vacancy are smaller, with magnitudes varying from 0.66 to 1.30 mT [15].

Tables Icon

Table 1. Comparison of 31P hyperfine interactions for the (Sb-VSn)0 trapped hole center and the isolated singly ionized Sn vacancy. Units are mT. For each defect, two 31P nuclei in one (P2S6) unit are participating (they are arbitrarily labeled nucleus 1 and nucleus 2).

An isothermal annealing study at 210 K was used to determine the stability of the (Sb-VSn)0 hole centers. The crystal was initially illuminated at 85 K with 510 nm light (obtained from a combined 150 W Xe lamp and band-pass filter). After removing the light, the intensities of the EPR spectra for the (Sb-VSn)0 and Sb2+ centers were recorded at 85 K. Then, with the sample still in the microwave cavity, the temperature was increased to 210 K and held there for 5 min, before returning to 85 K where the EPR spectra were recorded again. (Ramp rates up and down were 5 K/min.) This sequence was repeated for additional annealing times of 10, 15, and 30 min (the accumulated annealing time at 210 K was 60 min). This is similar to the process used in Reference [13] to determine activation energies. After 15 min of annealing at 210 K, 37% of the (Sb-VSn)0 signal remained and 92% of the Sb2+ signal remained. After 60 min of annealing, less than 20% of the (Sb-VSn)0 signal remained while 64% of the Sb2+ signal remained. Although both of the centers thermally decay at 210 K, the (Sb-VSn)0 hole centers are slightly less stable than the Sb2+ electron centers, and thus decay faster.

3.2 Spin-Hamiltonian parameters

The g matrix and the 121Sb hyperfine matrix describing the (Sb-VSn)0 center were obtained from the angular dependence of the EPR spectrum (measurements were made in three planes). These results are presented in Table 2. Corresponding principal values for the 123Sb hyperfine matrix can be extrapolated from the 121Sb results using the known nuclear magnetic moments. Results for the Sb2+ electron center [12] are also included in Table 2. For the (Sb-VSn)0 center, the 12 × 12 matrix form of the spin Hamiltonian (H = βS·g·B + I·A·S) was repeatedly diagonalized in a least-squares fitting process as the g and A matrix parameters were varied. It is important to note that the 121Sb hyperfine interactions are large and the hyperfine lines in the EPR spectrum are not equally spaced (see Fig. 1). Significant second-order shifts due to the large hyperfine principal values cause the spacing between adjacent Sb hyperfine lines to increase at higher magnetic field. In contrast, hyperfine interactions for the 31P nuclei are smaller, and the values presented in Table 1 are taken directly from separations between lines in the EPR spectra.

Tables Icon

Table 2. Spin-Hamiltonian parameters for the (Sb-VSn)0 trapped hole center and the Sb2+ trapped electron center. Principal-axis directions are expressed as (θ,ϕ) pairs. The polar angle θ is defined relative to the c axis and the azimuthal angle ϕ is defined relative to the a axis with positive rotation from a toward b in the plane perpendicular to c.

4. Discussion

As shown in several earlier investigations [7,10,11], the introduction of Sb3+ ions into SPS crystals is accompanied by a broad optical absorption appearing as a shoulder just below the optical band edge. In our sample, the peak of this poorly resolved optical absorption band is near 600 nm and its tail extends beyond 700 nm. The 510 nm light used in the present investigation is very close to the optical band edge. As expected, it transfers an electron from the valence band to the conduction band (at low temperature, the electrons and holes generated in this process are then trapped at the Sb-related donors and acceptors). Because 633 nm is well below the band edge, this light transfers an electron from the valence band to the Sb donor (without involving the conduction band). The hole left in the valence band is quickly trapped at the Sb-related acceptor. Thus, as we observe for 633 nm illumination at low temperature, this valence-band-to-donor transition leaves a trapped hole at the acceptor and a trapped electron at the donor.

Figures 2(a) and 2(b) show an isolated Sb3+ ion and an Sb3+ ion with an adjacent Sn vacancy, respectively. These schematic representations describe the Sb-related structures present in an SPS crystal before any illumination. The Sb3+ ions readily replace Sn2+ ions (the same bonding scheme is preserved because both ions have the 5s2 electron configuration). In the SPS crystal, the effective charge (relative to the regular unperturbed crystal) is different for these two Sb-related defects. During an illumination, the isolated Sb3+ ions in Fig. 2(a), with an effective charge of + 1, will trap photoinduced electrons whereas the Sb-VSn centers in Fig. 2(b), with an effective charge of −1, will trap photoinduced holes. Prior to the present study, Sb was expected to act only as an electron trap in SPS [12]. Now, the possibility of forming defect complexes involving adjacent donors and acceptors allows a donor such as Sb3+ to be part of a hole trap.

 figure: Fig. 2

Fig. 2 Models of Sb-related defects in an SPS crystal (before illumination), (a) Isolated Sb3+ ion with no nearby defects. During illumination, this defect attracts an electron and forms a trapped-electron center. (b) Sb3+ ion with a neighboring Sn vacancy. During illumination, this complex attracts a hole and forms a trapped-hole center.

Download Full Size | PDF

Our assignment of the new trapped-hole EPR spectrum in Fig. 1 (red stick diagram) to a Sn vacancy with an adjacent Sb3+ ion is based primarily on a comparison to the previously reported EPR spectrum from isolated singly ionized Sn vacancies [14]. These spectra have two features in common that provide support for this assignment. First, the two 31P hyperfine interactions are similar for the two defects (see Table 1). Second, hyperfine interactions with the nucleus at the neighboring cation site are similar (a 117Sn or 119Sn nucleus for the isolated Sn vacancy and a 121Sb or 123Sb nucleus for the (Sb-VSn)0 center. This second common feature is especially important. Approximately 22.5% of the unpaired spin density was estimated to be on the adjacent Sn ion in the singly ionized Sn vacancy [14], and the parameters in Table 2 suggest that a similar amount of spin density is on the Sb ion in the (Sb-VSn)0 center [12]. For both of these Sn-vacancy-associated defects, the observed hyperfine interactions with the neighboring cation are a result of hybridized sp2 (or sp2d5) combinations of S and Sn (or Sb) atomic orbitals [19]. More generally, there is hybridization between Sn (or Sb) atomic orbitals and P2S6 molecular orbitals. Future density-functional-theory (DFT) calculations will clarify the roles of covalency, electron lone pairs, and stereoactivity in the electronic structure of these holelike point defects, and also determine if the slightly smaller Sb3+ ion is displaced “off-center” at the Sn2+ site.

5. Summary

Electron paramagnetic resonance (EPR) is used to identify a new Sb-related defect in SPS. This defect consists of a charge compensating Sn vacancy next to a Sb3+ ion. During an illumination, this complex traps a hole and becomes EPR-active, i.e., a paramagnetic (Sb-VSn)0 center forms that is easily monitored at low temperature. Hyperfine structure in the spectra suggest strong similarities between the electronic ground state of the (Sb-VSn)0 center and the isolated singly ionized Sn vacancy [14]. When doping SPS with Sb, the relative concentrations of isolated Sb3+ ions and Sn-vacancy-associated Sb3+ ions may vary from crystal to crystal, depending on how many of the Sn vacancies (ranging from all to none) are located next to Sb3+ ions. Controlling factors include heating profiles during growth and possible post-growth heat treatments. Thus, not all Sb-doped SPS crystals are expected to have the same photorefractive response.

Acknowledgments

Dr. John Luginsland at the Air Force Office of Scientific Research (AFOSR Project 16RT0053) supported the work at the Air Force Institute of Technology, and the Science and Technology Center in Ukraine and the European Office of Aerospace Research and Development (STCU/EOARD Project P438a) supported the work at Uzhgorod National University. Views expressed in this paper are those of the authors and do not necessarily reflect the official policy or position of the Air Force, the Department of Defense, or the United States Government.

References and links

1. S. G. Odoulov, A. N. Shumelyuk, U. Hellwig, R. A. Rupp, A. A. Grabar, and I. M. Stoyka, “Photorefraction in tin hypothiodiphosphate in the near infrared,” J. Opt. Soc. Am. B 13(10), 2352–2360 (1996). [CrossRef]  

2. A. A. Grabar, I. V. Kedyk, M. I. Gurzan, I. M. Stoika, A. A. Molnar, and Y. M. Vysochanskii, “Enhanced photorefractive properties of modified Sn2P2S6,” Opt. Commun. 188(1–4), 187–194 (2001). [CrossRef]  

3. M. Jazbinšek, D. Haertle, G. Montemezzani, P. Günter, A. A. Grabar, I. M. Stoika, and Y. M. Vysochanskii, “Wavelength dependence of visible and near-infrared photorefraction and phase conjugation in Sn2P2S6,” J. Opt. Soc. Am. 22(11), 2459–2467 (2005). [CrossRef]  

4. M. Imlau, V. Dieckmann, H. Badorreck, and A. Shumelyuk, “Tin hypothiodiphosphate: nonlinear response in the sub-100 fs time domain,” Opt. Mater. Express 1(5), 953–961 (2011). [CrossRef]  

5. Y. Skrypka, A. Shumelyuk, S. Odoulov, S. Basun, and D. Evans, “Light-induced absorption and optical sensitizing of Sn2P2S6:Sb,” Opt. Commun. 356, 208–211 (2015). [CrossRef]  

6. A. Grabar, P. Mathey, and G. Gadret, “Manipulation of fast light using photorefractive beam fanning,” J. Opt. Soc. Am. B 31(5), 980–986 (2014). [CrossRef]  

7. P. Mathey, G. Gadret, A. Grabar, I. Stoika, and Y. Vysochanskii, “Photorefractive and photochromic effects in Sn2P2S6 at various temperatures,” Opt. Commun. 300, 90–95 (2013). [CrossRef]  

8. A. Grabar, P. Mathey, and R. Iegorov, “Coherent semilinear oscillator with Sn2P2S6:Sb crystals,” Appl. Phys. B 105(4), 813–819 (2011). [CrossRef]  

9. D. R. Evans, A. Shuymelyuk, G. Cook, and S. Odoulov, “Secondary photorefractive centers in Sn2P2S6:Sb crystals,” Opt. Lett. 36(4), 454–456 (2011). [CrossRef]   [PubMed]  

10. I. V. Kedyk, P. Mathey, G. Gadret, A. A. Grabar, K. V. Fedyo, I. M. Stoika, I. P. Prits, and Y. M. Vysochanskii, “Investigation of the dielectric, optical and photorefractive properties of Sb-doped Sn2P2S6 crystals,” Appl. Phys. B 92(4), 549–554 (2008). [CrossRef]  

11. T. Bach, M. Jazbinšek, G. Montemezzani, P. Günter, A. A. Grabar, and Y. M. Vysochanskii, “Tailoring of infrared photorefractive properties of Sn2P2S6 crystals by Te and Sb doping,” J. Opt. Soc. Am. B 24(7), 1535–1541 (2007). [CrossRef]  

12. A. T. Brant, L. E. Halliburton, S. A. Basun, A. A. Grabar, S. G. Odoulov, A. Shumelyuk, N. C. Giles, and D. R. Evans, “Photoinduced EPR study of Sb2+ ions in photorefractive Sn2P2S6 crystals,” Phys. Rev. B 86(13), 134109 (2012). [CrossRef]  

13. S. A. Basun, L. E. Halliburton, and D. R. Evans are preparing a manuscript entitled “Hyperbolic decay of photo-generated Sb2+ ions in Sn2P2S6 crystals detected with a novel electron paramagnetic resonance technique.”

14. E. M. Golden, S. A. Basun, D. R. Evans, A. A. Grabar, I. M. Stoica, N. C. Giles, and L. E. Halliburton, “Sn vacancies in photorefractive Sn2P2S6 crystals: An electron paramagnetic resonance study of an optically active hole trap,” J. Appl. Phys. 120(13), 133101 (2016). [CrossRef]  

15. E. M. Golden, S. A. Basun, A. A. Grabar, I. M. Stoika, N. C. Giles, D. R. Evans, and L. E. Halliburton, “Sulfur vacancies in photorefractive Sn2P2S6 crystals,” J. Appl. Phys. 116(24), 244107 (2014). [CrossRef]  

16. A. T. Brant, L. E. Halliburton, N. C. Giles, S. A. Basun, A. A. Grabar, and D. R. Evans, “Intrinsic small polarons (Sn3+ ions) in photorefractive Sn2P2S6 crystals,” J. Phys. Condens. Matter 25(20), 205501 (2013). [CrossRef]   [PubMed]  

17. G. Dittmar and H. Schäfer, “Crystal structure of Sn2P2S6,” Z. Naturforsch. C 29b(5–6), 312–317 (1974).

18. K. Kuepper, B. Schneider, V. Caciuc, M. Neumann, A. V. Postnikov, A. Ruediger, A. A. Grabar, and Y. M. Vysochanskii, “Electronic structure of Sn2P2S6,” Phys. Rev. B 67(11), 115101 (2003). [CrossRef]  

19. K. Glukhov, K. Fedyo, J. Banys, and Y. Vysochanskii, “Electronic structure and phase transition in ferroelectic Sn2P2S6 crystal,” Int. J. Mol. Sci. 13(12), 14356–14384 (2012). [CrossRef]   [PubMed]  

Cited By

Optica participates in Crossref's Cited-By Linking service. Citing articles from Optica Publishing Group journals and other participating publishers are listed here.

Alert me when this article is cited.


Figures (2)

Fig. 1
Fig. 1 EPR spectrum from an SPS crystal taken at 42 K with the magnetic field along the a axis. Blue stick diagrams identify lines from the Sb2+ ions (trapped electrons) and red stick diagrams identify lines from (Sb-VSn)0 complexes (trapped holes). Hyperfine lines from the 121Sb and 123Sb nuclei are present for both defects, as well as 31P hyperfine triplets for the (Sb-VSn)0 complexes.
Fig. 2
Fig. 2 Models of Sb-related defects in an SPS crystal (before illumination), (a) Isolated Sb3+ ion with no nearby defects. During illumination, this defect attracts an electron and forms a trapped-electron center. (b) Sb3+ ion with a neighboring Sn vacancy. During illumination, this complex attracts a hole and forms a trapped-hole center.

Tables (2)

Tables Icon

Table 1 Comparison of 31P hyperfine interactions for the (Sb-VSn)0 trapped hole center and the isolated singly ionized Sn vacancy. Units are mT. For each defect, two 31P nuclei in one (P2S6) unit are participating (they are arbitrarily labeled nucleus 1 and nucleus 2).

Tables Icon

Table 2 Spin-Hamiltonian parameters for the (Sb-VSn)0 trapped hole center and the Sb2+ trapped electron center. Principal-axis directions are expressed as (θ,ϕ) pairs. The polar angle θ is defined relative to the c axis and the azimuthal angle ϕ is defined relative to the a axis with positive rotation from a toward b in the plane perpendicular to c.

Select as filters


Select Topics Cancel
© Copyright 2024 | Optica Publishing Group. All rights reserved, including rights for text and data mining and training of artificial technologies or similar technologies.