You are on page 1of 7

jou.rm.

alof
ggclOar

ELSEVIER

Journal of Nuclear Materials 247 (1997) 315-321

Study on the hydrogen solubility in zirconium alloys


Shinsuke Yamanaka *, Masanobu Miyake, Masahiro Katsura
Department of Nuclear Engineering, Graduate School of Engineering, Osaka University, Yamadaoka 2-1, Suita, Osaka 565, Japan

Abstract
Pure zirconium and zirconium alloys of Zircaloy-2 (1.47 wt% Sn, 0.14 wt% Fe, 0.11 wt% Cr, 0.05 wt% Ni), Zircaloy-4
(1.52 wt% Sn, 0.21 wt% Fe, 0.11 wt% Cr) and Zr-1 wt% Nb alloy and Zr(O) alloys were selected as specimens. The
hydrogen solubility measurement was performed in the temperature range 500-1050C at a hydrogen pressure below 10 4
Pa. The hydrogen solubilities were different among alloys and the phase relationships were affected by the alloying element.
The partial thermodynamic quantities of hydrogen in the alloys were derived from the experimental data and the effect of the
alloying element was discussed. 1997 Elsevier Science B.V.

1. Introduction

2. Experimental

Zirconium alloys such as Zircaloy have been widely


used as the cladding materials of light water reactors. In
recent years, hydrogen behavior in the cladding has been
watched with keen notice because of high burn up of
nuclear fuel required from the economic point of view.
The renewed interest in hydrogen embrittlement of the
cladding promotes reassessment of the zirconium-hydrogen system.
The hydrogen solubility and the phase diagram for the
zirconium-hydrogen binary system have extensively been
studied [1-8] and reviewed [9-11]. The phase boundary
between the alpha phase and the hydride phase was examined for zirconium alloys below 500C [12,13]. However,
there exists insufficient data for the hydrogen solubility in
zirconium alloys at higher temperatures and a little information is available for the effect of gaseous contaminants
(mainly H 2 0 and 0 2) on the hydrogen solubility [14]. The
hydrogen solubility and the phase diagram are important
factors for the practical uses of zirconium alloys in light
water reactors under accidental as well as normal operating
conditions. From this point of view, the hydrogen solubility in substitutional and interstitial zirconium alloys has
been therefore examined in the present study.

Zircaloy-2, Zircaloy-4 and Zr-1 wt% Nb, the compositions of which are shown in Table 1, were chosen as
substitutional alloys and Zr(O) alloys with an oxygen
content of 0.025-0.02 O / Z r were chosen as interstitial
alloys. Hydrogen gas (initial purity, 99.999%) was further
purified by passing through a hot titanium bed and used
for solubility measurement.
The hydrogen solubility was measured in the temperature range 500-1050C at a hydrogen pressure below 104
Pa using a modified Sieverts' apparatus.

* Corresponding author. Tel.: +81-6 789 7887; fax: +81-6


873 3696; e-mail: yamanaka@nucl.eng.osaka-u.ac.jp.

3. Results and discussion


3.1. The effect of substitutional element

Figs. 1-8 show the equilibrium hydrogen pressure PH2


(Pa)-hydrogen concentration, C n ( H / Z r atom ratio),
isotherms for pure zirconium, Zircaloy-2, Zircaloy-4 and
Z r - I wt% Nb. The following phases appear: hcp a phase,
bcc [3 phase and fcc &
The hydrogen solubility in pure zirconium is known to
obey Sieverts' law in low hydrogen concentration region.
For zirconium alloys, Sieverts' law also holds between the
hydrogen pressure and the hydrogen concentration in both
the a phase and 13 phase regions, as shown in Figs. 1-8.
In our previous study [15], a comparison of Sieverts'

0022-3115/97/$17.00 1997 Elsevier Science B.V. All rights reserved.


PH S0022-3115(97)00101-3

316

S. Yamanaka et al. /Journal of Nuclear Materials 247 (1997) 315-321

Table 1
Compositions of zirconium alloys used in the present study
Zr a

Zircaloy-2

Zircaloy-4

1.47 wt%
0.14 wt%
0.11 wt%
0.05 wt%
0.125 wt%
bal.

wt%
0,21 wt%
0.11 wt%
< 35 ppm
0,125 wt%
bal.

(%)

104

Zr- 1Nb

- a o. o

875

. ,-a o o. o

Sn

Fe
Cr
Ni
Nb
O
Zr

74 ppm
< 50 ppm
bal.

530 ppm
0.97 wt%
0.143 wt%
bal.

t~O,o

" "-a * ~

950

~oaoo

975

ca

10

-,-aS~

1000

,~7~o
~oo

10
constant and the enthalpy of solution has been made
among zirconium alloys. In the c~ phase region (below
850C), Zircaloy-4 had the largest solubility among the
alloys and pure zirconium. The enthalpies of solution for
the alloys were higher than that for pure zirconium. In the
[3 phase region (above 950C), the hydrogen solubilities in
the alloys were smaller than that in pure zirconium. The
enthalpies of solution for the alloys were appreciably
higher than that for pure zirconium.
At higher hydrogen concentration, there exist discontinuities in the slope of isotherms below 850C for pure
zirconium, as evidenced by Fig. 1. The discontinuities
indicate the phase boundaries: c~ phase/[3 phase, [3
p h a s e / 5 phase and a p h a s e / ~ phase. Above 900C, the
slope of the isotherm is found from Fig. 2 to vary gradually with the hydrogen concentration, which suggests the
existence of a single [3 phase. There is not a large discrep-

10

0.01
0.1
C . (H/Zr)

(*c)
600
Q

650

_ ~.

600

,I~o

ancy between our results of the phase boundaries and the


assessed phase diagrams in the literature [9-11]. The
temperature of the eutectoid reaction: [ 3 Z r = a Z r +
5ZrH 2 .~ appears to be slightly higher than the 550C
reported in the literature [9-11].
The discontinuities in the slopes of the isotherms for
zirconium alloys were observed at the hydrogen concentra-

ol

1050

Fig. 2, Equilibrium hydrogen pressure Puz-hydrogen concentration, C H, isotherms for pure Zr. (Temperature: 875-1050C.)

.'?~.

550

,,~o
le-t~o

4
50O

102

lo 3

700
-aa

o oeooo:~

_\,,# oO

o=

a @ o

II,K

700

o-

8 qS) ,~

-102

....*-"

750

750

10

,_
o--

800

850

_ ao
~a o o

oo
o

cq

a0mnu

~ua

1
0.1

,,,,,d

850

"

'a" ,

1
......

....

0.01 0.1
C n(H/Zr)

IIHI

Fig. 1. Equilibrium hydrogen pressure PH2-hydrogen concentration, C H, isotherms for pure Zr. (Temperature: 500-850C.)

o
o o

4A-

-A

-ae
o_o
o_

o
u

~5 10

-A

e-
800

("4

.o

,,-a~o,a

1025

650

" oo * o

925

10 3

Total impurity concentration < 210 ppm.

a m o.

~xuo*O

1 03

9O0

1.52

u
oA

0.1

0.01 0.1
C H(H/Zr)

Fig. 3. Equilibrium hydrogen pressure P.2-hydrogen concentration, C H, isotherms for Zircaloy-2. (Temperature: 600-850C.)

317

S. Yamanaka et al. / Journal of Nuclear Materials 247 (1997) 315-321

104

10 4
900

mO

,~

900

,o

=o ~

a moaa

mOmoa 4P

925
925

10 3

*omo

103

950

950

. q ~
'e,%o

975

cq

10 2

975
o

IO00

"c'o~

1025

t'N

102

.~o~,

1000

~o~

~a~
1025

~'~

o
1050

10

10

1050

o,~
~oa

1
0.01

0.01

0.1

0.1

C - (H/Zr)

C .(H/Zr)
Fig. 4. Equilibrium hydrogen pressure PHz-hydrogen concentration, C a, isotherms for Zircaloy-2. (Temperature: 900-1050C.)

Fig. 6. Equilibrium hydrogen pressure Pna-hydrogen concentration, C a, isotherms for Zircaloy-4. (Temperature: 900-1050C.)

tion above 0.04H/Zr. No plateau pressure was found in


the isotherms for zirconium alloys. The regions above
0 . 0 4 H / Z r appear to correspond to o~ + [3 phase region.
Fig. 9 shows the experimental results obtained for the
phase boundaries between the ~ phase region and the

+ f3 phase region for the zirconium alloys ( O : Zry-2,


zx: Zry-4, O: Z r - l N b ) , together with the zirconium-hydrogen system (the dashed line and []). The dotted line
indicates the phase boundaries of the Z r - H system [10]
and the broken line indicates the reported values for
zirconium alloys [12,13]. The phase boundary slightly

104

('C)

1 04

~0

600

"a

650

(*C)

.-:

650

10

700

700

-"-102

750

750

10

o
a

~
a

io

S00
--OR

a a

800
t'-I

" o

850

Cq

850

10

o
o

~o
_ -a o
ao

a=

0.1
. . . . . . . .

0.01

1,,,,,I

0.1

0.1

,,,,,d

C H(H/Zr)
Fig. 5. Equilibrium hydrogen pressure PH2-hydrogen concentration, C u , isotherms for Zircaloy-4. (Temperature: 600-850C.)

i i iiHII

......

i , iliill

0.01
0.1
C .(H/Zr)

Fig. 7. Equilibrium hydrogen pressure P,%-hydrogen concentration, C a , isotherms for Zr-1 wt% Nb. (Temperature: 600-850C.)

318

S. Yamanaka et al./ Journal of Nuclear Materials 247 (1997) 315-321

CC)

omo

A~

900
.olo

~e

800

925

10 3

,olo ~ e

950

~I~AAe

/ " 0 Clb\
\

cv~

g~

975
o

10 2

\
\
02x t2a

700

1000

,~t,

I
1025

10

1050

*'i~

~qD
't~

I
I

~zx

600

0.01

, , ,,,,,I

, , ,,,,,

0.1

,/

.....J "

/..,.L/-

500

C H( H / Z r )
Fig. 8. Equilibrium hydrogen pressure PH2-hydrogen concentration, CH, isotherms for Zr-1 wt% Nb. (Temperature: 900lO5OOC.)

/;;,"

400

O.02

O. 04

O.06

O. 08

0.1

C (H/Zr)
differs among alloys and the addition of alloying elements
such as Sn, Fe, Cr, Ni and Nb appears to increase the
stability of the a phase. The effect of each alloying on
phase stability has not been clarified in the present study.
Our data appear to be consistent with low temperature data
sets for zirconium alloys [12,13]. Though the isotherms for
the alloys shown in Figs. 1-8 suggest that the phase
regions of 13, 13, + ~ and 8 may exist in the systems, the
phase boundaries at higher hydrogen concentrations were
not able to be determined in the present study.
The hydrogen solubility obviously changed with the
kind of alloy. Assuming that hydrogen atoms have access
to tetrahedral interstitial sites in a zirconium metal lattice,
we applied the following solubility equation [16] to the
experimental data to analyze the hydrogen solubility in the
zirconium alloys:

Zr
0

Zry-2 Zry-4 Z r - l N b
O

Zr

Zry, Zr-Nb

Reported value

Reported value

............................

Fig. 9. Phase boundaries between ~ and ~ + 13 for pure Zr and Z r


alloys. Experimental values: - o - :
Zr, O: Zry-2, A: Zry-4, O:
Zr-lNb.
Reported values: - . - , Z r [ 1 0 ] , - - - , zirconium alloys
[12,13].

-220
PureZr
[]

"

"5

~_~ -260

Zry2
O
Zr 4

~
-O--O--O--O-O - - % ~ ,

-o-o--o-~o

"~\

z,lNu

o~-\.~

~'e -'~',

In( C H T T / 4 / / ( /3 -- C H ) PI/22AH 2)

= - ( H H - E d ) / k T + SH/k.
In this equation, H H and S H are the partial molar enthalpy
and the partial molar excess entropy of hydrogen ,referring
to the standard state of hydrogen atoms at rest in a
vacuum, E d is one half of the dissociation energy of the
hydrogen molecule, k the Boltzmann constant and An2
the value related to partition functions of hydrogen gas.
The value o f / 3 is the number of interstitial sites of a given
kind available to hydrogen per zirconium atom. For the 13
phase, /3 = 6 and for the a phase, /3 = 2. In our previous
study [15], the values of partial molar thermodynamic

-300

0.01

0.1
CH(H/Zr)

Fig. 10. Change in the partial molar enthalpy H n value with the
hydrogen concentration C H.

S. Yamanaka et al. /Journal of Nuclear Materials 247 (1997) 315-321


quantities were estimated for the or phase of the alloys and
found to be constant irrespective of the hydrogen concentration. The partial molar quantities for the ct phase differed among the alloys. The results for the 13 phase
obtained in the present study are indicated in Figs. 10 and
11. As can be seen in Figs. 10 and 11, the partial molar
enthalpy and the partial molar excess entropy are constant
below 0.02 H / Z r and decrease with increasing hydrogen
concentration. The values of the partial molar quantities
vary with the kind of alloy and there is a marked difference in the concentration dependence among the alloys.
For both the ct and 13 phases, both the partial molar
enthalpy and excess entropy for dilute solution regions
increase with the addition of an alloying element. For the
ct phase, the magnitude of the changes in the partial molar
quantities is slightly larger for Zircaloy-4 than for the other
alloys, while for the 13 phase, it is larger for Z r - N b than
for the other alloys. Hydrogen-hydrogen interaction appears to cause the reduction in partial molar quantities for
the 13 phase and the magnitude of the interaction seems to
strongly depend on the kind of alloy. The positive changes
in the partial molar quantities with addition of alloying
elements cannot be explained by vibrational or dilatometric contribution. Further analysis is required to understand
the effect of each alloying element on the hydrogen solubility in zirconium.

319

Temperature (%)
1050

-5..4

1000

950

/
-5.6

A,v.**

or./3

.//z]//
//~/"

I"7~"
,/
,'~a /

--5.6
.../

-6.0

2 t';~

-6.2

.,'

-6~

7.4

76

28

B,O

I041T

Fig.

12.

Temperature dependence

~, *

DIll

0
*

8,2

8./*

(K -1 )

o f Sieve~s' constant K H for

J3Zr(O) alloys. (a) Pure Zr, (b) 0.0025 O/Zr, (c) 0.0058 O/Zr,
(d) 0.0102 O/Zr, (e) 0.0154 O/Zr and (f) 0.0203 O/Zr.

3.2. The effect of interstitial element


In our previous work [16], thermodynamic studies have
been performed on the hydrogen solubility in the hcp
ct Zr(O) alloy. The hydrogen solubility for the hcp c~ phase
first increased with the oxygen content and then decreased

40

PureZr

Zry2
0
Z~4

-0-0--0-00--0
--0-0--0-0 --0~ .
~Xk

Zrl Nb

20--~-~-~'~ ~'~\'\

%
I

0.01
0.,1
C~H/Zr)
Fig. 11. Change in the partial molar excess entropy SH value with
the hydrogen concentration CH.

at higher oxygen contents. The presence of interstitial


oxygen varied the enthalpy of solution.
All the solubility data for the bcc 13Zr(O) closely obey
Sieverts' law. The temperature dependence of Sieverts'
constant obtained for the 13Zr(O) alloys is shown in Fig.
12. It is found from Fig. 12 that linear relationships
between the natural logarithm of Sieverts' constant and the
reciprocal temperature hold for 13Zr(O) alloys with oxygen
contents below 0.01 O / Z r . There is no linear relationship
for alloys with higher oxygen contents. This suggests that
the Zr(O) alloys below 0.01 O / Z r is the single 13 phase
and the Zr(O) alloys with higher oxygen contents are in
the two phase region of ct + 13. The solubility of hydrogen
first decreases with the oxygen content and then increases
slightly. This trend for 13Zr(O) differs from that for ctZr(O).
The enthalpy of solution first decreases and then increases
as the oxygen content increases.
In order to clarify the influence of interstitial oxygen,
the solubility data for ctZr(O) alloys were analyzed by a
dilute solution model proposed for the ctZr(O) alloy in our
previous study [17]. The changes in the solubility and in
the enthalpy of solution with addition of interstitial oxygen
were discussed in terms of the difference in partial molar
enthalpy and excess entropy between pure metal and alloy,
which are defined as, ~ H = H H - H and ~ S = S H S
where H H and S H are the partial molar enthalpy and the
partial molar excess entropy of hydrogen in 13Zr(O) and
H and S are the values for pure 13 zirconium. Figs. 13

320

S. Yamanaka et al. / Journal of Nuclear Materials 247 (1997) 315-321

and 14 show the changes in 8 H and 8S values with the


composition of the alloy. As shown in Figs. 13 and 14, the
BH and BS first decrease with the oxygen content and then
markedly increase at higher oxygen contents. The change
in the solubility in [3Zr(O) alloys is attributable to a net
result from the variation in 8 H and 8S. The effects of
interstitial oxygen on partial molar quantities are significantly different from bcc [3Zr(O) to hcp aZr(O).
On the assumption that the enthalpy and entropy difference BH and BS were given in the form of 8 H = BH ~ +
8 H d + 8 H ~ and 8S = ~S ~ + 8S d + ~S c where 8 H v and
8S v are the differences in vibrational contribution between
the alloy and pure zirconium, 8 H d and 8S d the lattice
dilatation terms and 8 H ~ and 8S c the configurational
terms, we analyzed the changes in 5 H and ~S with the
oxygen content in [3Zr(O). The results for the analysis are
shown in Figs. 13 and 14. Detailed methods for the
analysis were published elsewhere [17].
As shown in Figs. 13 and 14, the ~ H v is small and
negative and 8S ~ is small and positive. The values of 8 H ~
and 8S v obtained for [3Zr(O) are larger than those for
o~Zr(O). As evidenced by Figs 13 and 14, the volume
expansion due to interstitial oxygen results in a negative
enthalpy change and the value of 8S d is negligibly small.
There is no marked difference in the volume effect between [3Zr(O) and oLZr(O). The configurational enthalpy

. . . . . . . . .

. . . . . . . . .

4~

. . . . . . . . .

L,

.-

.i

.f

-1

O.OOS

. . . . . . . . . . . . . . . . . . . . . . . . . .

0.01

":"':
0.01S

Composition [O/Zr]

term 8 H c increases with the oxygen content, as shown in


Fig. 13. Fig. 14 indicates that 8S c shows a negative value
and decreases with the oxygen content. The trend in 8 H c
and ~S c for 13Zr(O) is almost the same as c~Zr(O).
Although the theoretical curves shown by solid lines in
Figs. 13 and 14 demonstrate the tendency for the changes
in the experimental ~ H and 8S values with oxygen addition into f3Zr(O), the absolute values of 8 H and ~S were
different from one another
Although we have attempted to account for the influence of interstitial oxygen on hydrogen solubility in zirconium, it is not yet fully clarified. To explain the hydrogen
solubility in 13Zr(O), we have to take into account other
contributions such as the dual site occupation of hydrogen
in the bcc lattice. The dual site occupation of hydrogen
appears to result in increases in ~H c and 8S c.

Fig. 14. Theoretical analysis of the change in the 6S value with


the oxygen content of /3Zr(O). A : experimental values.

"

-"

-1

4. Conclusion
-2

0.005

0.01

0.01 5

Composition [O/Zr]
Fig. 13. Theoretical analysis of the change in the 6H value with
the oxygen content of ~Zr(O). (2): experimental values.

The hydrogen solubility in pure zirconium and zirconium alloys of Zircaloy-2, Zircaloy-4 and Zr-1 wt% Nb
and Zr(O) alloys has been studied to elucidate the influence of substitutional and interstitial elements on hydrogen
solubility in zirconium. The solubility measurement was

S. Yamanaka et al. / Journal of Nuclear Materials 247 (1997) 315-321

carried out at the temperatures between 500-1050C at a


hydrogen pressure below 104 Pa. The hydrogen solubilities were different among zirconium alloys and the phase
relationships were affected by the alloying element. The
influence of the alloying element was discussed on the
basis of the thermodynamic properties of hydrogen in
zirconium alloys derived from the experimental data.

Acknowledgements
The present study has been performed under the auspices of the Japan Atomic Energy Research Institute.

References
[1] C.E. Ells, A.D. McQuillan, J. Inst. Met. 85 (1956) 89.
[2] K. Watanabe, J. Nucl. Mater. 136 (1985) 1.

321

[3] M. Someno, J. Jpn. Soc. Met. 24 (1955) 136.


[4] E.A. Gulbransen, K.F. Andrew, J. Met. 7 (1955) 136.
[5] M.W. Mallet, W.M. Albrecht, J. Electrochem. Soc. 104
(1957) 142.
[6] R. Rica, T.A. Giorgi, J. Phys. Chem. 71 (1967) 3627.
[7] M. Tada, Y.C. Haung, Titanium-Zirconium 19 (1971) 260.
[8] M. Nagasaka, T. Yamashina, J. Less Common Met. 45
(1976) 53.
[9] E. Zuzek, Surf. Coating Technol. 28 (1986) 323.
[10] E. Zuzek, Bull. Alloy Phase Diagrams 11 (1990) 385.
[11] W.-E. Wang, D.R. Olander, J. Am. Ceram. Soc. 78 (1995)
3323.
[12] D. Hardie, W.H. Erickson, J. Nucl. Mater. 13 (1964) 254.
[13] J.J. Kearns, J. Nucl. Mater. 22 (1967) 292.
[14] M. Moalen, D.R. Olander, J. Nucl. Mater. 178 (1991) 161.
[15] S. Yamanaka, Y. Kashiwara, M. Miyake, J. Alloys Compounds 231 (1995) 730.
[16] S. Yamanaka, T. Tanaka, M. Miyake, J. Nucl. Mater. 167
(1989) 231.
[17] S. Yamanaka, M. Miyake, J. Nucl. Mater. 201 (1993) 134.

You might also like