Open Journal of Physical Chemistry, 2011, 1, 104-108
doi:10.4236/ojpc.2011.13014 Published Online November 2011 (http://www.SciRP.org/journal/ojpc)
Copyright © 2011 SciRes. OJPC
Atomistic Simulations of Formation of
Elementary Zr-I Systems
Matthew L. Rossi, Christopher D. Taylor
Materials Scie nce an d Tech nol o gy , MST-6, Los Alamos Na t ional La b, Los Alamos, USA
E-mail: mrossi@lanl.gov
Received August 6, 2011; revised September 8, 2011; accepted October 14, 2011
Abstract
We report results of simulations on the formation of simple zirconium iodide molecules. Previous work by
Wimmer et al. [1] explored the relationship between iodine and a zirconium surface. We investigate the re-
action schemes through atomistic simulations to better understand the nature of Zr-I interactions through iso-
lated molecules. The computed energy values of varying Zr-I systems suggests a strong binding mechanism
between zirconium and iodine, and offer predictions of likely reaction products. The computed results pre-
dict condensation of volatile ZrI4 with ZrI2 to form Zr2I6.
Keywords: Nuclear Chemistry, Quantum Chemistry, Computational Modelling
1. Introduction
Considerable amounts of research have been performed
regarding zirconium and zirconium-based alloys in the
past few decades, due to usage in cladding for light water
nuclear reactors (LWR).Stress-corrosion cracking (SCC)
of the cladding has been observed [2] over the course of
fuel-cycle lifetimes in nuclear reactors; the exact initia-
tion of which, however, remains yet unclear [3,4]. Much
research has been done regarding iodine-influenced SCC,
(I-SCC), which regards iodine as a strong accelerating
factor in the corrosion of the zirconium cladding [5-7].
In LWRs, the uranium fuel pellets are enclosed within
a zirconium-based alloy, such as Zircaloy-4 or Zirlo.
These materials are chosen due to the very low cross-
section of zirconium with respect to thermal neutrons.
However, these materials can undergo significant corro-
sion due to stress cracking and fission product reactivity,
in addition to other sources such as radiation or hydrogen
embrittlement. The effect of iodine, a fission product of
uranium, can affect the system in several ways. The first
is direct chemical interaction and changes to the basal
layer morphology characteristics due to bonding; the
second, a physical localized stress induced via iodine
adsorption. It is likely that I-SCC is caused by a combi-
nation of both effects on the exposed Zr surface, and
investigation is required to better understand the initia-
tion and nucleation mechanics of I-SCC. Physical stress
on the system can be caused by gas pressure of the fis-
sion products on the cladding walls. The internal fuel rod
pressures can range from 1 - 14 MPa during standard
operating conditions, up to 1 GPa in contained “bubbles”,
[8] with pressure increasing [9] with burn-up. The focus
of this research is the chemical interactions between zir-
conium and iodine.
The Zr cladding is expected to react strongly with the
released iodine during fission; iodine is indeed used to
purify Zr metal in a filament method [10]. Under this
scheme, iodine is used such that:
Zr+2I2 ZrI4 (1)
2I I
2 (2)
with the combination of the impure Zr and I2 gas
forming volatile ZrI4, which then deposits on a filament,
and can decompose under the equilibrium of (1), result-
ing in pure zirconium. This same mechanism is sus-
pected to be a means of I-SCC in Zr cladding [11]. For-
mation of zirconium iodides, however, can occur in mul-
tiple stoichiometries. There are structures provided in the
Inorganic Crystal Structure Database (ICSD) for ZrI2
[12], ZrI3 [13], in addition to ZrI4 [14]. One must con-
sider each of these molecular crystals in the effort to un-
derstand I-SCC, as it is likely that as iodine gas evolves
from the fuel pellet, the surface concentration of iodine
increases, and the surface Zr atoms undergo changes in
crystal morphology to reflect the current ZrIx stoich-
iometry. Shown in Equation (2), evolution of free iodine
atoms occurs during fission, which can combine to form
I2 gas, in cooperative equilibrium with (1).
105
M. L. ROSSI ET AL.
The formation of zirconium iodides are important due
to observations seen in previous research [15-17], which
can lead to pit formation and nucleation [5,11], followed
by material failures. A suspected mechanism of pit for-
mation is due to weakening of Zr-Zr bonds [7] at and
around the site of iodine adsorption. Once these bonds
are sufficiently weakened, it is likely that the zirconium
surface undergoes a period of iodine aggregation, where
the material proceeds from ZrxI1-x to a locally saturated
state approaching stoichiometric ZrI4. While this sort of
mechanism may not describe the bulk surface, such ag-
gregation is indeed likely to occur at or near grain
boundaries, where Zr-Zr bonding is significantly weaker
than along a pristine surface.
The research contained within this paper investigates
the agglomeration of iodine on zirconium atoms in iso-
lated systems.
2. Methodology
In an effort to understand the mechanistics of Zr-I inter-
actions, simple ZrIx molecular species were constructed
and energy structure calculations were performed. These
molecular species were chosen due to their correspond-
ing stoichiometries as given by the ICSD structures
[12-14]. In addition, mapping the potential energy of
dissociation for Zr2I6 was performed. ZrI4 volatilizes at
relatively low temperature [7] and can evolve from an
iodine-saturated Zr surface during reactor operation.
Therefore, understanding the iodine aggregation on the
Zr surface can provide insight to pitting and cracking [11]
which occur during cladding failure.
It is notable that Zr-I species undergo changes to spin
state due to bonding. ZrI2 is expected to exist in the
singlet state, while ZrI3 is in the doublet state and ZrI4 is
a singlet. The dimerized system, Zr2I6, is once again a
singlet. This can introduce problematic convergence in
systems due to crossover from triplet-to-singlet states as
the isolated materials approach. The reaction schemes
are as follows (shown in Figure 1):
ZrI2 + ZrI4 Zr2I6 (3a)
2ZrI3 Zr2I6 (3b)
Note that in (3b), the reactants are in the triplet state;
while this triplet state is caused by the presence of two
doublets, a singlet state is also possible. The reverse re-
action of 3a and 3b are disproportionation (dis) and
comproportionation (com), respectively.
Additional reaction schemes were investigated to
study the propagation of iodizing zirconium. These
schemes were used to simulate the means by which Zr
metal results in the volatile ZrI4 species, which has been
observed to cause pitting and corrosion [5,11]. Elemen-
Figure 1: Reaction schematic for the decomposition of Zr2I6
into disproportionation (upper) and comproportionation
(lower) components. Electronic structure calculations pre-
dict formation of Zr2I6 from both the D and C materials.
tary formations of higher-valency iodides were formed
under the following reactions:
Zr + I2 ZrI2 (4a)
ZrI2 + I2 ZrI4 (4b)
2ZrI2 + I2 2ZrI3 (5a)
2ZrI3 + I2 2ZrI4 (5b)
All calculations were performed with the GAMESS
[18,19] software package using the hybrid B3LYP
[20-22] functional with the SBKJC [23-25] basis set.
3. Results and Discussion
ZrI2 was calculated in both the singlet and triplet state,
based on the ground state electron configuration of Zr
(3F2). In the case of ZrI2, the singlet state is preferred by
a value of 1.24 eV. This is consistent with bonding to the
d-orbital and the presence of a non-bonding fully occu-
pied Zr s-orbital.
By computation of (3a), we observe that the equilib-
rium favors the product, Zr2I6, and does not undergo dis-
proportionation. The energy change of the forward reac-
tion (3a) is 3.32 eV. The reaction pathway between the
products and reactants in (3b) is also exothermic, with
computed value of 1.79 eV. Therefore, in the event of
dissociation, the overwhelmingly favored product would
be comproportionation to ZrI3, with an energetic favor-
ing of 1.54 eV relative to the disproportionation prod-
ucts.
Formation of Zr2I6 from component materials, as
shown in (3a,b), is of primary interest. As originally re-
ported by Busol [26], there are two possible mechanisms
in ZrIx formation:
3ZrI4 + Zr 4ZrI3 (7a)
ZrI4 + Zr 2ZrI2 (7b)
Busol [26] determined that the rate limited step in ZrI3
formation was not I2 concentration, but rather, the diffu-
Copyright © 2011 SciRes. OJPC
M. L. ROSSI ET AL.
106
sion rate at which I2 can pass through surface ZrIx layers
to reach pristine Zr. Observations from their work have
shown that formation of ZrI3 is the primary reaction,
with little or no formation of ZrI2. This is consistent with
the energetic barrier we observe to the formation of ZrI2
from Zr2I6, which at high temperature the latter can de-
compose. This, however, does not preclude volatilization
of ZrI4, which can form when the exposed Zr surface is
sufficiently saturated with iodine, such that the dispro-
portionation reaction (Equation (3a)) need not occur to
yield ZrI4.
In a system containing a pair of ZrI3 molecules, sepa-
rated at 6Å beyond the equilibrium separation of Zr at-
oms in Zr2I6 (from here on, denoted eq + 6Å), the pre-
ferred electronic state is to remain as two doublets which
combine for an overall triplet state, energetically favored
by 1.43 eV relative to that of the singlet state. This en-
ergy difference is evidence of errors within Hartree-Fock
(HF) methods, as two isolated ZrI3 molecules should
exhibit identical energies regardless of being in the sing-
let or triplet state. Upon inspection of the atomic charges,
it does not appear that any ionization of species occurred,
but that the ZrI3 molecules were essentially symmetric in
charge distribution. Examination of the molecular orbi-
tals (Figures 2-3) show the loss of degeneracy in the
isolated ZrI3-ZrI3 singlet state, as is expected. However,
the population of the non-degenerate HOMO shows
electron density primarily in the d-orbitals, suggesting a
delocalized, higher energy orbital occupation which does
not occur when the system is allowed to exist in the trip-
let state, as a sum of two doublets.
Higher level theory was implemented (MCSCF
[27-31]/SBKJC) to investigate this deviation, with resul-
tant energy between the singlet and triplet states showing
negligible (4.55 × 10–5 eV) difference. Regardless of
choosing restricted or restricted open-shell Hartree-Fock,
HF-based methods fail to properly describe two isolated
doublets, as these methods do not properly handle the
multi-configurational spin states involved. Therefore, the
energy difference predicted between singlet and triplet
isolated ZrI3 pairs is an artifact of Hartree-Fock theory.
However, MCSCF is significantly more computationally
expensive than DFT, and therefore is useful for the in-
vestigation of the difference between singlet and triplet
states, but impractical for the high volume of calculations
required to determine the potential energy curve mapped
in Figure 4. The energy of the dissociated ((eq) + 6Å)
disproportionation products was also computed to make
a comparison between the dis and com lowest-energy
products relative to the DFT results shown in Figure 4,
with an energy difference of 2.05 eV favoring the com-
proportionation products. This energy is slightly larger
than DFT-predicted result of 1.54 eV, both favoring
Figure 2. Molecular orbital diagram of ZrI3-ZrI3 singlet
system at eq + 6.0Å separation. Black indicates occupied
orbitals, grey indicates virtual (unoccupie d) orbitals.
Figure 3. Molecular orbital diagram of ZrI3-ZrI3 triplet
system at eq + 6.0Å separation. Black indicates occupied
orbitals, grey indicates virtual (unoccupie d) orbitals.
formation of 2ZrI3 in the dissociation of Zr2I6.
Mapping of the potential energy curve for the reaction
coordinate (Figure 4) for combination/dissociation of
Zr2I6 to the comproportionation and disproportionation
products shows that at approximately Zr-equilibrium (eq)
+ 1Å separation, the Zr2I6 system undergoes a spin trans-
formation from the singlet Zr2I6 to a roughly associated
triplet ZrI3-ZrI3 pair.
The combination schemes (3a, 3b, shown in Figure 4)
reveal that in the region between the equilibrium Zr-Zr
separation of Zr2I6 to a separation of ~2Å, both the dis
and com behave identically, reaching a stable, loosely
bound Zr2I6 when the iodines are allowed to relax during
a constrained Zr-Zr distance optimization. However, as
the separation increases, the asymptotic limit is reached
Copyright © 2011 SciRes. OJPC
107
M. L. ROSSI ET AL.
Figure 4. Energy curves for disproportionation (diamonds)
and comproportionation in the singlet (squares) and triplet
(triangles) states relative to displacement from the equilib-
rium position. The dashed line indicates the suspected path
which includes conversion from the isolated triplet (two
doublet) state of ZrI3 to combination resulting in the singlet
Zr2I6.
for the dis and com products, respectively, which differ
slightly in energy. It can be noted, however, that due to
the preference for the singlet-triplet transition of the co m
products at a separation of ~eq + 1Å, the com reaction
will instead follow the dashed (low) path (Figure 4).
The reactions show in Equations (4)-(5) yielded exo-
thermic results, consistent with increasing iodine coor-
dination to the Zr central atoms. The associated energies
of these reactions is shown in Table 1. The computed
enthalpies of Equations (4b) and (5a) are lower the en-
ergy change associated with 5b. The data suggests that
there is a preference to the formation of ZrI3, rather than
ZrI4, again, supported by the report by Busol [26]
4. Conclusions
It is apparent that aggregation of ZrIx species will form
larger systems as shown in structures from the ICSD
[12-14]. The choice of studying Zr2I6 was based on the
analogous structure of bulk ZrI2 [12], which is not
stoichiometrically identical, but geometrically similar as
the system size increases. (The excess of iodine in the
stoichiometry is due to terminal iodine capping.) For ex-
ample, the next largest structure, Zr3I8, with each Zr atom
having two bridging iodines, and maintaining a tetrahe-
dral geometry of the Zr cores, consistent with the bulk
ZrI2 structure. Such aggregation supports the idea that as
iodine exposure increases, similar to what occurs within
a nuclear reactor, the iodine will bind to the Zr cladding,
causing changes to the surface morphology of the clad-
ding material, and enabling the growth of poly
crystalline ZrI2. It is expected that upon reaching some
Table 1. Energy change table for reactions 4-5.
Reaction Reactants Products ΔE (eV)
4a Zr + I2 ZrI2 –4.59
4b ZrI2 + I2 ZrI4 –6.78
5a 2ZrI2 + I2 2ZrI3 –9.50
5b 2ZrI3 + I2 2ZrI4 –5.24
critical limit, the surface iodine saturation will cause
morphological changes to the Zr surface, resulting in a
mixture of ZrIx species, and ultimately to stoichiometric
ZrI4. Upon reaching iodine surface saturation, it is then
possible that molecular ZrI4(g) volatilization will be pos-
sible, resulting in pitting, and ultimately, corrosion of the
cladding.
5. Acknowledgements
This research was supported by the Consortium for Ad-
vanced Simulation of Light Water Reactors (www.
casl.gov), an Energy Innovation Hub (http://www.en-
ergy.gov/hubs) for Modeling and Simulation of Nuclear
Reactors under U.S. Department of Energy Contract No.
DE-AC05-00OR22725. The Los Alamos National Labo-
ratory is operated by Los Alamos National Security LLC
for the National Nuclear Security Administration of the
U.S. Department of Energy under contract DE-AC52-
06NA25396.
6. References
[1] E. Wimmer, R. Najafabadi, et al., “Ab Initio Calculations
for Industrial Materials Engineering: Successes and Cha-
llenges,” Journal of Physics: Condensed Matter, Vol. 22,
No. 38, 2010, p. 384215.
doi:10.1088/0953-8984/22/38/384215
[2] S. A. Nikulin and A. B. Rozhnov, “Corrosion Cracking of
Zirconium Cladding Tubes (A Review). I. Methods of
Study and Mechanisms of Fracture,” Metal Science and
Heat Treatment, Vol. 47, No. 1-2, 2005, pp. 71-79.
doi:10.1007/s11041-005-0034-2
[3] S. B. Goryachev, A. R. Gritsuk, et al., “Iodine Induced
SCC of Zr Alloys at Constant Strain Rate,” Journal of
Nuclear Materials, Vol. 199, No. 1, 1992, pp. 50-60.
doi:10.1016/0022-3115(92)90439-R
[4] V. V. Likhanskii and L. V. Matweev, “The Development
of the Crack Growth Model in Zirconium Claddings in
Iodine Environment,” Nuclear Engineering and Design,
Vol. 213, No. 2-3, 2002. pp. 133-140.
doi:10.1016/S0029-5493(01)00516-7
[5] A. Serres, L. Fournier, et al., “The Effect of Iodine
Content and Specimen Orientation on Stress Corrosion
Crack Growth Rate in Zircaloy-4,” Corrosion Science,
Vol. 52, No. 6, 2010, pp. 2001-2009.
doi:10.1016/j.corsci.2010.02.008
Copyright © 2011 SciRes. OJPC
M. L. ROSSI ET AL.
Copyright © 2011 SciRes. OJPC
108
[6] I. Schuster and C. Lemaignan, “Influence of Texture on
Iodine-Induced Stress Corrosion Cracking of Zircaloy-4
Cladding Tubes,” Journal of Nuclear Materials, Vol. 189,
No. 2, 1992, pp. 157-166.
doi:10.1016/0022-3115(92)90528-S
[7] P. P. S. Sidky, “Iodine Stress Corrosion Cracking of
Zircaloy Reactor Cladding: Iodine Chemistry (a Re-
view),” Journal of Nuclear Materials, Vol. 256, No. 1,
1998, pp. 1-17. doi:10.1016/S0022-3115(98)00044-0
[8] F. Lemoine, “High Burnup Fuel Behavior Related to
Fission Gas Effects under Reactivity Initiated Accidents
(RIA) Conditions,” Journal of Nuclear Materials, Vol.
248, No. 1, 1997, pp. 238-248.
doi:10.1016/S0022-3115(97)00157-8
[9] P. P. Rudling, R. Adamson, et al., “High Burnup Fuel
Issues,” Nuclear Engineering Technology, Vol. 40, 2008,
1-8.
[10] K. D. Sinel’nikov, F. I. Busol and G. I. Stepanova,
“Problem of the Iodine Method of Purification of Zir-
conium,” Atomic Energy, Vol. 4, 1958. pp. 221-227.
[11] S. Y. Park, et al., “Crack Initiation and Propagation
Behavior of Zirconium Cladding under an Environment
of Iodine-Induced Stress Corrosion,” Metals and Ma-
terials International, Vol. 13, No. 2, 2007, pp. 155-163.
doi:10.1007/BF03027567
[12] D. H. Guthrie and J. D. Corbett, “Synthesis and Structure
of an Infinite-Chain Form of Zr I2 (alpha),” JSSC, Vol.
37, 1981, pp. 256-263.
[13] A. Lachgar, D. S. Dudis and J. D. Corbett, “Revision of
the Structure of Zirconium Triiodide. The Presence of
Metal Dimers,” Inorganic Chemistry, Vol. 29, No. 12,
1990, pp. 2242-2246. doi:10.1021/ic00337a013
[14] B. Krebs, G. Henkel and M. Dartmann, “Kristallstruktur
von Zirkoniumtetrajodid Zr I4. Ein Neuer A B4
Strukturtypp,” Acta Crystallographica Section B, Vol. 35,
1979, pp. 274-278. doi:10.1107/S0567740879003344
[15] S. B. Farina, G. S. Duffo and J. R. Galvele, “Localized
Corrosion of Zirconium and Zircaloy-4 in Iodine Alco-
holic Solutions,” LAAR, Vol. 32, 2002, pp. 295-298.
[16] S. B. Farina, G. S. Duffo and J. R. Galvele, “Stress
Corrosion Cracking of Zircaloy-4 in Halide Solutions.
Effect of Temperature,” Materials Research, Vol. 5, No.
2, 2002, pp. 107-112.
doi:10.1590/S1516-14392002000200004
[17] P. P. Jacques, F. Lefebvre and C. Lemaignan, “Defor-
mation-Corrosion Interactions for Zr Alloys during I-
SCC Crack Initiation. Part I: Chemical Contributions,”
Journal of Nuclear Materials, Vol. 264, No. 3, 1999, pp.
239-248. doi:10.1016/S0022-3115(98)00501-7
[18] M. W. Schmidt, K. K. Baldridge, et al., “General Atomic
and Molecular Electronic Structure System,” Journal of
Computational Chemistry, Vol. 14, No. 11, 1993, pp.
1347-1363. doi:10.1002/jcc.540141112
[19] M. S. Gordon and M. W. Schmidt, “Advances in
Elec- tronic Structure Theory: GAMESS a Decade
Later,” In: C. E. Dykstra, G. Frenking, et al., Eds.,
Theory and App- lications of Computational
Chemistry, the First Forty Years, Elsevier,
Amsterdam, 2005, pp. 1347-1363.
doi:10.1016/B978-044451719-7/50084-6
[20] A. D. Becke, “Density-Functional Thermochemis-
try. III. The Role of Exact Exchang e, ” Journal of
Chemical Physics, Vol. 98, 1993, pp. 5648-5642.
doi:10.1063/1.464913
[21] P. P. J. Stephens, F. J. Devlin, et al., “Ab Initio
Calculation of Vibrational Absorption and Circular
Dichroism Spectra Using Density Functional Force
Fields,” The Journal of Physical Chemistry, Vol. 98, No.
45, 1994, pp. 11623- 11627. doi:10.1021/j100096a001
[22] R. H. Hertwig and W. Koch, “On the Parameterization of
the Local Correlation Functional. What Is Becke-
3-LYP?” Chemical Physics Letters, Vol. 268, No. 5-6,
1997, pp. 345-351. doi:10.1016/S0009-2614(97)00207-8
[23] W. J. Stevens, H. Basch and M. Krauss, “Compact
Effective Potentials and Efficient Shared-Exponent Basis
Sets for the First- and Second-Row Atoms,” Journal of
Chemical Physics, Vol. 81, No. 12, 1984, pp. 6026-6033.
doi:10.1063/1.447604
[24] W. J. Stevens, M. Krauss, H. Basch and P. P. G. Jasien,
“Relavtivistic Compact Effective Potentials and Efficient,
Shared-Exponent Basis-Sets for the 3rd-Row, 4th-Row
and 5th-Row Atoms,” Canadian Journal of Chemistry,
Vol. 70, No .2, 1992, pp. 612-630. doi:10.1139/v92-085
[25] T. R. Cundari and W. J. Stevens,Effective Core
Potential Methods for the Lanthanides,” Journal of
Chemical Physics, Vol. 98, No. 7, 1993, pp. 5555-5565.
doi:10.1063/1.464902
[26] F. I. Busol, “Kinetics of the Reduction of ZrI4 Vapors by
Metallic Zirconium,” Russian Journal of Physical Che-
mistry, Vol. 33, 1959, pp. 799-807.
[27] M. W. Schmidt and M. S. Gordon, “The Construction and
Interpretation of MCSCF Wavefunctions,” Annual
Review of Physical Chemistry, Vol. 29, 1998, pp. 233-
266. doi:10.1146/annurev.physchem.49.1.233
[28] B. O. Roos, “The Multiconfiguration SCF Method,” In: S.
W. Diercksen, Ed., Methods in Computational Molecular
Physics, D. Reidel, Dordrecht, 1983, pp. 161-187.
[29] J. Olsen, D. L. Yeager and P. P. Jorgensen, “Optimization
and Characterization of a MCSCF State,” Advances in
Chemical Physics, Vol. 54, 1983, pp. 1-176.
doi:10.1002/9780470142783.ch1
[30] H. -J. Werner, “Matrix Formulated Direct MCSCF and
Multiconfiguration Reference CI Methods,” Advances in
Chemical Physics, Vol. 69, 1987, pp. 1-62.
doi:10.1002/9780470142943.ch1
[31] R. Shepard, “The MCSCF Method,” Advances in Che-
mical Physics, Vol. 69, 1987, pp. 63-200.
doi:10.1002/9780470142943.ch2