+ All Categories
Home > Documents > Author's personal copy · Author's personal copy Structural development in Ge-rich Ge S glasses Y....

Author's personal copy · Author's personal copy Structural development in Ge-rich Ge S glasses Y....

Date post: 24-Sep-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
6
This article appeared in a journal published by Elsevier. The attached copy is furnished to the author for internal non-commercial research and education use, including for instruction at the authors institution and sharing with colleagues. Other uses, including reproduction and distribution, or selling or licensing copies, or posting to personal, institutional or third party websites are prohibited. In most cases authors are permitted to post their version of the article (e.g. in Word or Tex form) to their personal website or institutional repository. Authors requiring further information regarding Elsevier’s archiving and manuscript policies are encouraged to visit: http://www.elsevier.com/copyright
Transcript
Page 1: Author's personal copy · Author's personal copy Structural development in Ge-rich Ge S glasses Y. Sakaguchia,*, D.A. Tenneb, M. Mitkova a aDepartment of Electrical and Computer Engineering,

This article appeared in a journal published by Elsevier. The attachedcopy is furnished to the author for internal non-commercial researchand education use, including for instruction at the authors institution

and sharing with colleagues.

Other uses, including reproduction and distribution, or selling orlicensing copies, or posting to personal, institutional or third party

websites are prohibited.

In most cases authors are permitted to post their version of thearticle (e.g. in Word or Tex form) to their personal website orinstitutional repository. Authors requiring further information

regarding Elsevier’s archiving and manuscript policies areencouraged to visit:

http://www.elsevier.com/copyright

Page 2: Author's personal copy · Author's personal copy Structural development in Ge-rich Ge S glasses Y. Sakaguchia,*, D.A. Tenneb, M. Mitkova a aDepartment of Electrical and Computer Engineering,

Author's personal copy

Structural development in Ge-rich Ge–S glasses

Y. Sakaguchi a,*, D.A. Tenne b, M. Mitkova a

a Department of Electrical and Computer Engineering, Boise State University, Boise, Idaho 83725-2075, United Statesb Department of Physics, Boise State University, Boise, Idaho 83725-1570, United States

a r t i c l e i n f o

Article history:Available online 28 July 2009

PACS:61.43.FS64.75.Yz78.30.Ly

Keywords:Raman scatteringChalcogenidesRaman spectroscopy

a b s t r a c t

The Raman spectra of Ge–S glasses in the Ge-rich region from Ge 33% to 46% have been investigated inorder to know the structural development of the network glasses. From the detailed curve fits, we havefound that there is an unassigned peak at 410 cm�1 and it becomes larger with increasing Ge composi-tion. To clarify the structural origin of the peak, we virtually constructed the atomic arrangement ofthe glassy state starting from the crystalline state through the liquid state and changed the compositiongradually depleting the medium in sulfur. From the consideration of the structural modeling and theatomic orbital theory, we suggest that single Ge–S chain is a probable structural origin of the peak.

� 2009 Elsevier B.V. All rights reserved.

1. Introduction

Ge-chalcogenide (Ge-X: X = S, Se) glasses are known as typicalnetwork glasses and are of a great interest because of their appli-cation potential. To utilize their attractive properties, it is quiteimportant to know the exact structure of the materials. In the pastfew decades there has been an extensive effort to investigate thestructure of the glassy Ge-chalcogenides. Results of several mea-surements have shown that a structural transition occurs withincreasing Ge composition [1,2]. According to the results ofMössbauer spectroscopy [2], at high Ge composition there arethree structural phases, A, B and C, in these glasses. A phase hassimilar structure to the high temperature crystalline phase ofGeX2, which consists of tetrahedral Ge(X1/2)4 units. B phase con-sists of the ethane-like (X1/2)3Ge–Ge(X1/2)3 units. C phase has adouble layer structure like crystalline(c-) GeX. The relative pres-ence of these phases changes with increasing Ge composition. Inrecent years, the interest was mainly focused on the region aroundthe critical composition of the rigidity percolation threshold,where the number of the constraints per atom is equal to the de-gree of freedom. In the composition region, a steep first-order-liketransition, from very flexible floppy phase to rigid phase, is ob-served. The composition in Ge–Se system is Ge 23% [3]. In addition,in very recent years, great attention has been paid to a narrow re-gion near the critical composition, the ‘intermediate phase win-

dow’, where specific changes due to a self-organization occur[3,4]. For Ge–Se system, the region is between Ge 20% and 26%[3]. On the other hand, in the Ge-rich region from Ge 33% to43%, a much qualitative change is observed in the Raman spectra[1]. Indeed, it can be related to the structural transformations fromA to B and from B to C phase. However, it is still worth examiningthe details of the structural development in the composition re-gion because a quantitative analysis of Raman spectra has notbeen fully made yet. In this paper, we present the results of the de-tailed analysis of the Raman spectra of Ge–S glasses in the Ge-richregion, including our experimental data of thin films. On the basisof the results, we propose a unified model, which also explainsother previous experimental results. Using the concept of themodel, we finally provide perspective on the possibilities forapplications.

2. Experimental

Thin films of amorphous Ge46S54 with the thickness of 300 nmwere prepared by sputtering. The composition was measured bythe electron probe micro-analyzer in the system of the scanningelectron microscope (LEO 1430VP). Raman spectra of the films atroom temperature in a vacuum chamber were recorded using a Ra-man spectroscopic system of Horiba Jobin Yvon T64000, in back-scattering geometry. The 441.6 nm He–Cd laser line was used forexcitation. The laser power was changed from 5 to 73 mW andwe have not observed a spectral change with increasing the laserpower. Therefore, we adopted the spectrum obtained with73 mW, which had the best signal to noise ratio.

0022-3093/$ - see front matter � 2009 Elsevier B.V. All rights reserved.doi:10.1016/j.jnoncrysol.2009.04.064

* Corresponding author. Present address: Quantum Beam Directorate, JapanAtomic Energy Agency, 2-4 Shirane, Shirakata, Tokai-mura, 319-1195, Japan. Tel.:+81 29 282 6769; fax: +81 29 284 3889.

E-mail address: [email protected] (Y. Sakaguchi).

Journal of Non-Crystalline Solids 355 (2009) 1792–1796

Contents lists available at ScienceDirect

Journal of Non-Crystalline Solids

journal homepage: www.elsevier .com/ locate / jnoncrysol

Page 3: Author's personal copy · Author's personal copy Structural development in Ge-rich Ge S glasses Y. Sakaguchia,*, D.A. Tenneb, M. Mitkova a aDepartment of Electrical and Computer Engineering,

Author's personal copy

3. Results

Fig. 1 shows the Raman spectra of Ge-rich Ge–S glasses from Ge33% to 46%. The three Raman spectra, (a)–(c), are from the resultsof Takebe et al. [5], which were measured for bulk samples. Thespectrum at Ge 46% was obtained using thin film prepared by sput-tering in the present study. The qualitative change in the spectra isconsistent with the results of Lucovsky et al. [1] and Kotsalas andRaptis [6]. A background component has been subtracted for thespectra at Ge 40% and 46%. We performed curve fits with Gaussiansto the spectra, assuming the presence of the peaks, which have al-ready been assigned. Among the peaks, the peak at 340 cm�1 isattributed to the symmetric breathing mode of S atoms atGe(S1/2)4 tetrahedron [7]. The intensity of the peak decreases withincreasing Ge composition. This suggests that the number of thetetrahedral units decreases with increasing Ge composition.According to the work by Boolchand et al. [2], the content of Aphase is 0% at Ge 36%. Although the content might not be perfectlyzero, it is supposed to be small and there can be other structuralorigin for the peak at 340 cm�1 in the spectrum at Ge 36%. Accord-ing to Lucovsky et al. [8], there is a vibrational frequency at340 cm�1 of the ethane-like units, whose Raman active frequenciesare at 240, 340 and 376 cm�1. The presence of the ethane-like unitsindicates the presence of Ge–Ge bonds, in other words, chemicaldisorder in the system. There is a variety of interpretations onthe peak at 370 cm�1. The asymmetric stretching mode of the tet-rahedral unit is located at 375 cm�1 [7,9]. The peak at 370 cm�1

was often referred to as a companion peak and it was interpretedas the stretching motion of the outrigger raft (OR) accompaniedby S–S bond, which was suggested by Bridenbaugh et al. [10]. Re-cently, the peak is often assigned as the vibration of S atoms on theedge-sharing double bonds [11,12]. The vibrational frequency ofthe ethane-like units is also located near 370 cm�1 [8]. Jacksonet al. point out from first-principles molecular-dynamics simula-tions that there are two peaks near 370 cm�1; the mode of theedge-sharing cluster (373 cm�1) and the mode of the ethane-like

cluster (366 cm�1) [13]. The peak at 430 cm�1 is regarded as thestretching mode of dimerized S atoms on the edge of the OR[14]. This is also a peak indicative of chemical disorder at a stoichi-ometric composition of Ge 33%. The peak decreases with increasingGe composition from 33% to 36%, and almost vanishes at Ge 40%.One can expect that the chance for creation of S–S bond declineswith decreasing S composition and the result is consistent withthe expectation. In 200–300 cm�1, we obtained two peaks locatedat 220 and 255 cm�1 from the curve fit. One of the vibrational fre-quencies of the ethane-like unit is located near 255 cm�1;240 cm�1 according to Lucovsky et al. [8] and 255 cm�1 accordingto Jackson et al. [13]. In the spectra of c-GeS, there are two inten-sive peaks at 212 and 238 cm�1, which are the modes of doublelayer structure [15]. The peak at 220 cm�1 is considered to origi-nate from these peaks. So far we made curve fits only using thepeaks which have already been assigned. However, in the Ge-richregion more than 35%, we could not fit the spectra without consid-ering a peak at 410 cm�1, which becomes larger with increasing Gecomposition. It is obvious that the peak at 410 cm�1 is not formeddue to a shift of the peak at 430 cm�1 because both peaks were re-quired to fit the spectrum at Ge 36%. It seems that the increase ofthe peak at 410 cm�1 is related to the increase of C phase. But thereis no peak with such high frequency in the spectra of c-GeS. Wewould expect that there exists a structural unit, which has a stron-ger bond than that in the double layer and gives such high fre-quency vibration. As far as we know, no one suggests such astructural unit.

4. Discussion

4.1. Structural model

In order to find out the structural origin of the vibrational modeat 410 cm�1, we have performed a virtual structural modeling.Fig. 2 shows how amorphous structure is formed from c-GeS2. Hightemperature crystalline phase of GeS2 consists of the tetrahedral

50 100 150 200 250 300 350 400 450 500

A = 0 %B = 15 %*

C = 75 %*

A = 0 %B = 60 %C = 40 %

A = 0 %B = 80 %C = 20 %

A = 73 %B = 27 %C = 0 %

(a)430

Ge33S67

340

370

(b)255

220

410Ge36S64

(c)Ge40S60

(d)Ge46S54

Scat

tere

d in

tens

ity

Raman shift (cm-1)

Fig. 1. Raman spectra of Ge–S glasses and the results of the curve fit. The dots show the experimental Raman spectra. The solid curves show the results of fitting, which arethe sum of the indicated peaks. The contents of A, B and C phase, which were obtained from Mössbauer spectroscopy [2], are also indicated. The values at Ge 46% wereestimated from the extrapolation of the results.

Y. Sakaguchi et al. / Journal of Non-Crystalline Solids 355 (2009) 1792–1796 1793

Page 4: Author's personal copy · Author's personal copy Structural development in Ge-rich Ge S glasses Y. Sakaguchia,*, D.A. Tenneb, M. Mitkova a aDepartment of Electrical and Computer Engineering,

Author's personal copy

units [16]. In the figure, there are two streams of Ge–S chains. Thestream can also be regarded as the sequence of corner-sharing tet-rahedral units. The edge-sharing tetrahedral units connect the twostreams. When c-GeS2 is heated and molten GeS2 is obtained fromthe high temperature crystalline phase, some Ge–S covalent bondswould break. If a bond break occurs as shown in the figure, twoneighboring S atoms become free and there is a chance to form anew S–S bond. This corresponds to the S–S dimmer on the edgeof the OR. After breaking of the Ge–S bonds, the other counterpartsof Ge atoms become also free. These atoms can bind each other andform a new Ge–Ge bond. As a result, the ethane-like unit is formed.Such structure formed in the liquid phase is supposed to be pre-served in the amorphous phase by quenching. This picture pro-vides an answer to the fundamental question, why the chemicaldisorder exists in the stoichiometric composition and in amor-phous phase.

To clarify the structure in Ge-rich Ge–S glasses, we subtracted Satoms from the structure of amorphous (a-) GeS2 as shown inFig. 3. Here we subtracted S atoms outside of the Ge–S chain andwe assumed that the Ge–S chain structure is preserved. S atomsare classified into three groups; 1st, 2nd and 3rd column, as shownin the figure. The S atoms in the 1st column are the compositionalelements of the Ge–S chain. The S atoms in the 2nd column are atthe edges of the corner-sharing tetrahedral units on the Ge–Schain, but not in the Ge–S chain. The S atoms in the 3rd columnare in the middle of the two Ge–S chains, and are outside of the2nd column. By subtracting S atoms, the ethane-like unit will loseone S atom and Ge atom will seek other S atom. If there is a neigh-boring S atom in the 2nd column, the Ge atom will bond with this Satom. If there is no neighboring S atom in the 2nd column, the Geatom will bond with a S atom in the 1st column. Through suchdynamics, we can expect that the interaction between the twostreams of Ge–S chain becomes stronger and that the two streamsapproach each other. Here, the ethane-like units play a role of glue.In fact, the composition variation of density indicates that there isconsiderable volume contraction in the region from Ge 33% to 45%[5,17]. The same tendency is also observed in Ge–Se system [18].We suggest that the large volume contraction in the region iscaused by such an increase of the interaction between the Ge–S(Se) chains.

4.2. Formation of layer structure

To know further about the structural nature of the Ge–S glassesclose to Ge 50%, the diffraction study is useful. Using the results ofthe neutron diffraction study for Ge–Se system by Salmon’s group[19,20], we obtained that the bond angles of Se–Ge–Se are 111.01o

in a-GeSe2, and 101.57o in liquid (l-) GeSe2, which are close to109.47�, while in liquid GeSe the bond angle is 94.82�, which isclose to 90�. Recent ab initio molecular-dynamics simulations byVan Roon et al. support these bond angles [21]. The change inthe bond angle from Ge 33% to 50% can be explained in terms ofthe atomic orbital theory. The outermost electronic configurationof Ge is (4s)2(4p)2 and that of Se is (4s)2(4p)4. In Se atom, two selectrons are low in energy and they do not participate in bonding.Two electrons among four p electrons are used to form covalentbond with other two atoms. Hence, Se atom makes twofold coordi-nation. The remaining two p electrons do not participate in bond-ing and form a pair called ‘lone-pair electrons’. For Ge atom, thereare two cases depending on the composition. When the Se compo-sition is large enough, a tetrahedral structure will be formed cen-tering a Ge atom. In this case, sp3 hybridization occurs in the Geatom by the promotion of one s electron to p level. The formationof four sp3 orbitals results in the four equivalent covalent bonds.Thus, the bond angle is close to 109.47�, which is the value for aperfect tetrahedron. When the Ge composition becomes 50%,twofold coordinated Ge–Se chains are formed, as depicted inFig. 4(a), as a result of the subtraction of Se atoms. In order to formtwofold coordination in Ge atom, two p electrons would partici-pate in bonding, leaving s electrons in bonding without forminghybridization. Since two p orbitals are perpendicular to each other,the bond angle of Se–Ge–Se is fixed to be 90�. In Fig. 4(b), we haveillustrated how a double layer structure can be formed from theGe–Se chains. The Ge–Se chains are alternatively aligned along b-axis at different height in c-axis. They are connected to each otherhaving a third bond. This is the way how a layer structure is formedfrom the chains and the layer becomes ‘double’. The formation ofthe third bond can also be explained in terms of the atomic orbitaltheory. In Se atom, two p electrons are used for covalent bonds andtwo p electrons remain as lone-pair electrons. In Ge atom, two pelectrons are used for covalent bonds and one p orbital is empty.

Fig. 2. A structural model, which explains how chemical disorder is formed in amorphous GeS2 from crystalline GeS2 through the liquid phase.

1794 Y. Sakaguchi et al. / Journal of Non-Crystalline Solids 355 (2009) 1792–1796

Page 5: Author's personal copy · Author's personal copy Structural development in Ge-rich Ge S glasses Y. Sakaguchia,*, D.A. Tenneb, M. Mitkova a aDepartment of Electrical and Computer Engineering,

Author's personal copy

Between the lone-pair orbital and the empty orbital, a coordinatebond can be formed as the possibility of the formation in chalco-genide glasses has been discussed by Ovshinsky and Adler [22].Applying this concept to studied system, we can expect that anew coordinate bond could be formed between a Ge atom and aSe atom, which are in different neighboring chains.

4.3. Origin of the peak at 410 cm�1

The important point in the third bond formation is that thebond formation is possible only for hetero-polar pair. Accordingto the neutron diffraction study, there exist homo-polar bonds inl-GeSe [19]. The homo-polar bonds should only exist in the

Fig. 3. A model of the structural development in Ge–S glasses from Ge 33%(a-GeS2) to Ge-rich region.

Fig. 4. (a) Atomic arrangement at Ge 50%. Se atoms are subtracted from the structure of GeSe2. (b) Formation of the double layer structure from Ge–Se chains. Color of thechain corresponds to that in (a). (c) Formation of the third bond (indicated by blue in (b)). (For interpretation of the references to colour in this figure legend, the reader isreferred to the web version of this article.)

Y. Sakaguchi et al. / Journal of Non-Crystalline Solids 355 (2009) 1792–1796 1795

Page 6: Author's personal copy · Author's personal copy Structural development in Ge-rich Ge S glasses Y. Sakaguchia,*, D.A. Tenneb, M. Mitkova a aDepartment of Electrical and Computer Engineering,

Author's personal copy

Ge–Se chain composed of covalent bonds. This leads to the situa-tion that some atoms in the chain face homo-polar atoms in aneighboring chain. At the portion, the atoms do not bond withthe faced atoms and the chains can fail to form double layer. Inthe region, the Ge–Se chain is left as a single chain without beinginvolved in the layer. The structure in the liquid might bequenched in amorphous phase. Although we dealt with Ge–Se sys-tem in the above discussion, the situation can be the same in Ge–Ssystem since the atomic organization and chemical bonding arevery similar. Therefore, such single chain in GeS system would bea possible structural origin of the peak at 410 cm�1.

5. Conclusion

We have made an analysis of Raman spectra of Ge-rich Ge–Sglasses and found that there is an unassigned peak, which is lo-cated at 410 cm�1, in the region from Ge 36% to 46%. In order toknow the structural origin of the peak, we considered the struc-tural development in the atomic arrangement starting from thecrystal structure of GeS2 and proposed a structural model of Ge-rich Ge–S glasses. The model explains why the ethane-like unitsappear in the glasses. The interaction between the two Ge–S chainsbecomes stronger with increasing Ge composition due to the roleof the ethane-like units. This is consistent with the data of the den-sity measurements. We have also demonstrated that a new type ofbond, which bridges the two Ge–S chains to form double layerstructure, can appear in Ge-rich region close to 50% using a conceptof atomic orbital theory. From these our considerations, we havespeculated that there is incomplete portion in the double layerstructure in Ge-rich Ge–S glasses close to 50% and proposed thestructural origin of the Raman peak at 410 cm�1 as single Ge–Schain. Although we could not give direct evidence in the presentstudy, such possibility will be examined by first-principles molec-ular-dynamics simulations. Our model emphasizes the uniquenessof amorphous structure and its way to self-organize. The concept ofour model can provide a useful guideline to find a new functional-ity in Ge–S glasses in Ge-rich region. The unique nature in theinteraction between Ge–S chains can affect optoelectronic re-sponse and photodiffusion of silver. In addition, the low dimen-sionality in the structure may be utilized in producing a newtype of nano-structures. It would be interesting to examine thesepossibilities in further studies.

Acknowledgements

The authors are grateful to C. Raptis and K. Nakamura forinvaluable discussion and to H. Takebe and P. Boolchand for per-mission to use their data for analysis. The authors also thank IdahoMicrofabrication Laboratory in Boise State University for technicalassistance in composition measurements. D. A. T. acknowledgessupports from DOE EPSCOR Grant No. DE-FG02-04ER46142. Y.S.acknowledges supports from IMI- NFG (NSF Grant No. DMR-0409588).

References

[1] G. Lucovsky, F.L. Galeener, R.C. Keezer, R.H. Geils, H.A. Six, Phys. Rev. B 10(1974) 5134.

[2] P. Boolchand, J. Grothaus, M. Tenhover, M.A. Hazle, R.K. Grasselli, Phys. Rev. B33 (1986) 5421.

[3] P. Boolchand, X. Feng, W.J. Bresser, J. Non-Cryst. Solids 293&295 (2001) 358;P. Boolchand, D.G. Georgiev, B. Goodman, J. Optoelectron. Adv. Mater. 3 (2001)703.

[4] M.F. Thorpe, D.J. Jacobs, M.V. Chubynsky, J.C. Phillips, J. Non-Cryst. Solids266&269 (2000) 859.

[5] H. Takebe, H. Maeda, K. Morinaga, J. Non-Cryst. Solids 291 (2001) 14.[6] I.P. Kotsalas, C. Raptis, Phys. Rev. B 64 (2001) 125210.[7] G. Lucovsky, J.P. deNeufville, F.L. Galeener, Phys. Rev. B 9 (1974) 1591.[8] G. Lucovsky, R.J. Nemanich, F.L. Galeener, in: W.E. Spear (Ed.), Proceedings of

the 7th International Conference on Amorphous and Liquid Semiconductors,Edinburgh, Scotland, G.G. Stevenson, Dundee, Scotland, 1977, p. 125.

[9] N. Kumagai, J. Shirafuji, Y. Inuishi, J. Phys. Soc. Jpn. 42 (1977) 1262.[10] P.M. Bridenbaugh, G.P. Espinosa, J.E. Griffiths, J.C. Phillips, J.P. Remeika, Phys.

Rev. B 20 (1979) 4140.[11] S. Sugai, Phys. Rev. B 35 (1987) 1345.[12] K. Murase, K. Inoue, O. Matsuda, in: Y. Sakurai, Y. Hamakawa, T. Masumoto, K.

Shirae, K. Suzuki (Eds.), Current Topics in Amorphous Materials: Science andTechnology, Elesevier, Amsterdam, 1993, p. 47.

[13] K. Jackson, A. Briley, S. Grossman, D.V. Porezag, M.R. Pederson, Phys. Rev. B 60(1999) R14985.

[14] K. Murase, T. Fukunaga, Y. Tanaka, K. Yakushiji, I. Yunoki, Physica 117B&118B(1983) 962.

[15] J.D. Wiley, W.J. Buckel, R.L. Schmidt, Phys. Rev. B 13 (1976) 2489.[16] Von G. Dittmar, H. Schäfer, Acta Crystallogr. B 31 (1975) 2060.[17] Y. Kawamoto, S. Tsuchihashi, J. Am. Ceram. Soc. 54 (1971) 131.[18] A. Feltz, H. Aust, A. Blayer, J. Non-Cryst. Solids 55 (1983) 179.[19] P.S. Salmon, I. Petri, J. Phys.: Condens. Matter 15 (2003) S1509.[20] I. Petri, P.S. Salmon, H.E. Fischer, J. Phys.: Condens. Matter 11 (1999) 7051.[21] F.H.M. van Roon, C. Massobrio, E. de Wolff, S.W. de Leeuw, J. Chem. Phys. 113

(2000) 5425.[22] S.R. Ovshinsky, D. Adler, Contemp. Phys. 19 (1978) 109.

1796 Y. Sakaguchi et al. / Journal of Non-Crystalline Solids 355 (2009) 1792–1796


Recommended