You are on page 1of 4

phys. stat. sol. (c) 4, No. 3, 1008– 1011 (2007) / DOI 10.1002/pssc.

200673704

Investigation into thallium sites and defects


in doped scintillation crystals
A. N. Blacklocks*, 1, A. V. Chadwick1, R. A. Jackson2, and K. B. Hutton3
1
Functional Materials Group, School of Physical Sciences, University of Kent, Canterbury,
Kent CT2 7NH, UK
2
Lennard-Jones Laboratories, School of Physical and Geographical Sciences, Keele University, Keele,
Staffs, ST5 5BG, UK
3
Hilger Crystals, Westwood, Margate, Kent CT9 4JL, UK

Received 29 June 2006, revised 29 August 2006, accepted 25 September 2006


Published online 9 March 2007

PACS 29.40.Mc, 61.72.Ji


Thallium doped caesium iodide, CsI(Tl), and sodium iodide, NaI(Tl) are two of the most efficient scintil-
lators developed and are already widely used for radiation detection and imaging applications. Their use
in fast imaging applications however has been hindered by a long lasting high level of afterglow – the
percentage of the luminescence pulse remaining a short time after excitation. Very little is known about
the point defects in these crystals, such as structure and concentrations, and the first step to understanding
the causes of the afterglow is to understand the nature of the defects responsible for the scintillation. In
this paper the local structure of the thallium activator ion has been investigated via EXAFS spectroscopy
and some basic intrinsic defects calculated using the General Utility Lattice Program (GULP).

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

1 Introduction In recent years there has been a large increase in demand for improved imaging sys-
tems for industrial, security and medical applications. One of the more established single crystal scintil-
lators in these fields is cadmium tungstate, CdWO4, but there are concerns with the safety and production
of this material with cadmium use being progressively banned [1]. These concerns have resulted in a
search for cheaper and safer alternatives for medical and freight imaging in particular. One problem is
that a large number of replacement scintillators suffer from a high level of long term afterglow – the
percentage of luminescence pulse remaining a short period after excitation. This delayed emission can
cause reduced contrast and image blurring in high speed x-ray imaging. Manufacturers of scanners are
now setting tight limits on afterglow levels and times with very few materials consistently able to meet
the criteria. Two potential replacements are thallium activated forms of caesium iodide, CsI(Tl) and
sodium iodide, NaI(Tl). These single crystals provide two of the highest photoelectron yields reported
when coupled to a Si-photodiode (CsI(Tl)) or a bi-alkali photomultiplier (NaI(Tl)). However, both crys-
tals generally have unaccepTable levels of afterglow, as high as 5% of the scintillation pulse after 2 ms
[2] depending on the strength and duration of the excitation pulse, although it can vary from sample to
sample and seems difficult to control in a systematic manner.
It is generally accepted that the afterglow is caused by trapping of the carriers that re-combine with the
excited defect centres. Contributors to the afterglow are therefore assumed to be impurities [3], intrinsic
lattice defects [2–4] and mechanically induced stress the crystals [5]. Only recently has a reasonable
mechanism for the scintillation process been agreed [2, 6] although the kinetics remain very complex
with as many as 12 distinct mechanisms involved in the luminescence of these crystals [7]. One problem

*
Corresponding author: e-mail: anb6@kent.ac.uk, Phone: +44-1227-823043, Fax: +44-1227-827558

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


phys. stat. sol. (c) 4, No. 3 (2007) 1009

preventing the full understanding of the causes of the afterglow is the lack of data on the basic point
defects in the crystals such as structures and concentrations. The main aim of this study is therefore to
identify point defects present in these systems.

2 Experimental

2.1 Materials Single crystal samples of CsI(Tl) and NaI(Tl), pure CsI, NaI and TlI were all obtained
from Hilger Crystals. All single crystals used for this work were grown from the melt using the Kyro-
polous method under a dry nitrogen atmosphere to reduce oxygen and water content.

2.2 EXAFS spectroscopy Measurements were made on station 16.5 at the CCLRC Daresbury Syn-
chrotron Radiation Source. The synchrotron is a 2GeV electron storage ring operating at circulating
currents between 150 and 250 mA. Thallium LIII edge EXAFS spectra were recorded at liquid nitrogen
temperature in fluorescence mode using a 30 element Ge solid-state detector with high count-rate elec-
tronics. Data was processed using the Daresbury EXAFS programmes EXCALIB and EXCURV98 [8]
and the Klementiev background subtraction software VIPER [9]. For each spectrum a theoretical fit was
made by adding shells of atoms around the central excited atom and least squares refining the radial
distance, RD, and the Debye-Waller type factors, 2σ2, which contain contributions from thermal disorder
and static variations in RD.

2.3 Defect calculations All calculations have been performed using the GULP code [10]. Interac-
tions between ions are represented by effective potentials, with ionic polarisability included via the shell
model.

Table 1 Comparison of experimental and calculated lattice parameters.

(i) NaI (ii) TlI


Parameter Exp. Calc. Diff. (%) Parameter Exp. Calc. Diff. (%)
3 3
Volume (Å ) 287.496 287.549039 0.02 Volume (Å ) 309.392656 308.477284 -0.30
a=b=c (Å) 6.6 6.600406 0.01 a (Å) 4.57 4.488824 -1.78
α=ß=γ (°) 90 90 0 b (Å) 12.92 13.213285 2.27
c (Å) 5.24 5.200915 -0.75
α=ß=γ (°) 90 90 0

Table 2 Interatomic potential parameters for CsI, NaI and TlI.

Interaction A (eV) ρ (Å) C (eVÅ6) Interaction k (eVÅ-2) Y(|e|)


Csshell – Ishell 4913 0.3814 480.06 Cscore – Csshell 244.84 −6.6337
Nashell – Ishell 6900 0.3038 51.79 Nacore – Nashell 292.43 −2.2510
Tlshell – Ishell 2800 0.3814 480.06 Tlcore – Tlshell 244.84 −6.6337
Ishell – Ishell 2532 0.4474 1015.26 Icore – Ishell 19.59 −3.0332

The CsI and NaI starting potentials were taken from Ref. [11], with the final NaI and TlI potentials used
produced by fitting to the crystal structures as shown above in Table 1. In the defect calculations, consis-
tent cut-off radii of 10 and 15 Å have been used for regions I and II(a) respectively.

www.pss-c.com © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim


1010 A. N. Blacklocks et al.: Thallium sites and defects in doped scintillation crystals

3 Results and discussion The Tl LIII edge EXAFS was firstly measured for thallium (I) iodide – the
raw material used for doping the CsI and NaI. A good fit was achieved by fitting to the crystallographic
data for orthorhombic TlI [12] giving confidence in the thallium and iodine phase shifts applied. Data
was also collected for NaI(Tl) with a nominal thallium doping level of 1000ppm with the EXAFS and
Fourier Transforms both shown below in Fig. 1. The thallium appears to be present on a sodium site with
a close fit to the face-centred cubic NaI model as shown below in Table 3. There is clearly a local expan-
sion around the central Tl+ compared to Na+ as would be expected from the differing radii (Tl+ 1.64 Å,
Na+ 1.16 Å). The apparent shell at ~2.5 Å in the Fourier Transform (Fig. 1) is an artefact caused by the
thallium phaseshifts.

Table 3 Shell model data for NaI(Tl) from both crystallographic data [13] and EXAFS fitting.

(i) Crystallographic data (ii) EXAFS

CN Atom RD (Å) CN Atom RD (Å) 2σ2 (Å2)


6 I 3.235 6 I 3.373 0.010
12 Na 4.575 12 Na 4.576 0.025
8 I 5.603 8 I 5.600 0.035
Ef = -0.739 R = 38.42

7
20
5
FT Magnitude

3 15
k3 χ(k)

1
10
-1 3 4 5 6 7 8 9
-3 5

-5
0
-7 0 2 4 6 8 10
-1
Wavenumber (Å ) Radial Distance (Å)

Fig. 1 Tl L-III edge EXAFS (left) and Fourier Transform (right) of NaI(Tl). Solid lines are experimental data, and
dashed lines are the theoretical fit.

Spectra were also recorded for CsI(Tl) samples, but due to poor data quality (very noisy above ~5 Å–1) a
poor fit is achieved to any model input resulting in a need for higher quality data before this material can
be fully analysed.

In Table 4 we list the calculated intrinsic defect formation energies, and also the migration energies in
the case of the vacancies. Also listed is the solution energy for substituting a Tl+ onto a Na+ or Cs+ site.
In this case a simple solution energy scheme is assumed:

TlI + MM → TlM + MI (where M= Cs, Na)


Esol = −Elatt (TlI) + Esubs (TlM) + Elatt (MI)

It is noted that our defect energies are generally in good agreement with those of Rowell and Sangster
[14] and Lynch [15]. Our data suggest that Schottky defects predominate with binding energies of 0.48
and 0.45 eV in CsI and NaI. Intrinsic defects involving interstitials in CsI appear unlikely when their
formation energies are compared to Schottky defects although Rowell does suggest that Frenkel defects

© 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.pss-c.com


phys. stat. sol. (c) 4, No. 3 (2007) 1011

could make significant contributions to conductivity at high temperature due to very low migration ener-
gies indicated by their calculations.

Table 4 Defect formation and migration energies (in eV) for CsI and NaI at 0 K. Values in brackets are
taken from Ref. [14]. Inset is a representation of the face-centred (A) and edge centred (B) interstitial site
positions used for CsI.

Defect CsI NaI


Formation Migration Formation Migration
Vacancy Cation 3.89 (4.09) 0.44 (0.55) 4.56 (4.38) 0.82 (0.85)
Anion 4.07 (4.01) 0.34 (0.39) 5.11 (5.00) 1.01 (1.04)
Schottky Bound 1.04 − 1.89 −
Unbound 1.52 (1.86) − 2.34 (2.34) −
Tl solution 0.30 − −0.06 −
Formation
"A" (0 ½ ½) "B" (½ 0 0)
Interstitial Cation 0.20 (0.04) −0.80 (−0.51) A
Anion 0.28 (0.90) 0.52 (0.76)
Frenkel Cation 4.09 3.27 (3.53)
Anion 4.35 4.59 (4.62) B

In CsI we see that anion migration is slightly more favourable whereas in NaI the smaller and less po-
larisable cation migration is favoured by the vacancy mechanism. The calculated solution energies for Tl
in CsI and NaI are 0.30 and −0.06 eV respectively. This suggests that Tl should substitute very easily
onto both Cs and Na sites, although it should be noted that assumptions are involved here such as infinite
dilution of defects and a temperature of 0 K. More detailed studies are also currently in progress involv-
ing calculating experimental defect energies from doped CsI single crystals via a.c. conductance meas-
urements and relating the defect structures to the afterglow phenomenon.

Acknowledgements We wish to thank Hilger Crystals for the award of an EPSRC CASE studentship to ANB and
also acknowledge support for this work from the Franco-British Interreg IIIA Programme (grant 31530).

References
[1] EU Directive 2002/95/EC, Off. J. Eur. Union L37, 19 (2003)
[2] C. Brecher, V.V. Nagarkar, V. Gaysinskiy, S.R. Miller, and A. Lempicki, Nucl. Instrum. Methods Phys. Res. A
537, 117 (2005).
[3] C. Greskovich and S. Duclos, Annu. Rev. Mater. Sci. 27, 69 (1997).
[4] M. Nikl, Meas. Sci. Technol. 17, R37 (2006).
[5] V. Gayshan, A. Boyarintsev, A.V. Gektin, and D.I. Zosim, Nucl. Instrum. Methods Phys. Res. A 505, 97
(2003).
[6] S. Zazubovich, Radiat. Meas. 33, 699 (2001).
[7] A. Lempicki, A.J. Wojtowicz, and C. Brecher, Inorganic Scintillators, in: S. R. Rotman (ed.) Wide Gap Lumi-
nescent Materials: Theory and Applications (Kluwer Academic Press, Norwell 1999), p. 235.
[8] N. Binsted, EXCURV98: CCLRC Daresbury Laboratory computer program (1998).
[9] K.V.Klementev, J. Phys. D: Appl. Phys. 34, 209 (2001).
[10] J.D. Gale, J. Chem. Soc., Faraday Trans. 93, 629 (1997).
[11] M.J.L. Sangster and R.M. Atwood, J. Phys. C: Solid State Phys. 11, 1541 (1978).
[12] G.A. Samara, L.C. Walters, and D.A. Northrop, J. Phys. Chem. Solids 28, 1875 (1967).
[13] H.E. Swanson and M. Mehmel, Natl. Bureau Stand. 539, 31 (1955).
[14] D.K. Rowell and M.J.L. Sangster, J. Phys. C: Solid State Phys. 14, 2909 (1981).
[15] D.W. Lynch, Phys. Rev. 118, 468 (1960).

www.pss-c.com © 2007 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

You might also like