Sei sulla pagina 1di 17

5.

Elastic Interactions of Hydrogen in the


Lattice of Iron Alloys
M.P. Puts
Whiteshell Nuclear Research Establishment
Atomic Energy of Canada Limited
Pinawa, Manitoba, Canada

TABLE OF C O N T E N T S

1.
2.

5.
6.

Introduction
General Formulation of Elastic Interaction
2.1 Volume Change and Distortion Field
2.2 Interaction Energy
2.3 The Diaelastic Polarizability
2.4 The Paraelastic Polarizability
Thermodynamics of Stressed Solids Containing Mobile
Interstitials
Application to Hydrogen in Iron
4.1 Interstitial Site
4.2 Partial Molal Volume and Strain Field Symmetry
4.3 Defect Moduli
Conclusions
References

1.

INTRODUCTION

3.
4.

The response of hydrogen to applied stress fields can be


described by an elastic interaction, i.e., a description of the
interaction in terms of elastic distortions. The usefulness of
this representation is limited to strains and stresses for which
linear elasticity is valid. This restricts the validity of the
elastic interaction model to small, externally applied loads and
to long-range interactions of hydrogen with other defects such
as dislocations, cracks and solute atoms.
Hydrogen in iron is a mobile interstitial solute. Within
the elastic model, it is represented as a source of elastic distortion occupying the interstices of the body-centered-cubic
(bcc) lattice. The formulation of the distortion field of a
point defect and its relation to external and internal crystal
volume changes are presented here, along with a derivation of
the elastic interaction energy. The thermodynamics of stressed
solids containing mobile interstitials and the effect of the
elastic interaction on the solubility of hydrogen in the lattice
are also discussed. A numerical evaluation of the elastic interaction requires knowledge of the magnitude and symmetry of the
strain field of hydrogen and the response of this strain field

to applied stresses. The experimental and theoretical evidence


providing information on these parameters is examined.
2.
2.1

GENERAL FORMULATION QF ELASTIC INTERACTION


Volume Change and Distortion Field

As stated in the introduction, hydrogen in iron is an interstitial point defect that can be represented either as an
internal source of strain (1) or stress (2). We can obtain an
expression for the interaction that is independent of the details
of the microscopic distortions near the defect, and it matters
little which representation of the defect is chosen.
However,
for a defect with a complex structure, such as an interstitial
atom, derivation of the interaction energy in terms of stress is
mathematically and conceptually most convenient, a point that
will become particularly evident when making comparisons with
lattice models. Therefore, a treatment is presented in terms of
the stress representation, and this is compared later with the
results based on the strain representation.
In the stress representation (for the most part, the treatment of Leibfried and Breuer (2) has been followed here), the
defect is simulated by an internal set of forces, f. These must
be such as to give vanishing total force and torque. The displacement field s_(r_) caused by these forces is given in terms of
the elastic Green's function by
S 1 Cr) = jdrGik(r-r')fk(r')

(1)

where r
~
r_'

= position vector at which the displacement is


calculated
= position vector at which the force is calculated
(jik(r_-rf) = static elastic Green's function
fk(r_f)
= k-th component of force at r'
i,k
= 1,2,3; Cartesian components

and, unless otherwise indicated, repeated indices imply summation.


Physically, the static Green's function is the displacement
field due to a unit force. The asymptotic behaviour of s(r) for
distances much larger than the region over which operates" is
obtained by expansion of G_ in powers of r', This~leads to a
series in powers of 1/r, with the leading term giving
s. (r) * - P , 3 G., (r)
i -nk n ik J

(2)

with the moment

nk

= l H r ' x- x -t f fl f r] ' t -

J-- n k -

>

nk

=F P

kn

l jfSl

where Xn' is a Cartesian component of r', and Pn^ is called the


dipole or double force tensor of the force distribution f(r').
The dipole tensor completely describes the macroscopic response
of the crystal to the distortions produced by a single point
defect.
The above description is very useful for comparison with a
lattice model. In this case, the displacement field of the ,
point defect can be simulated by the application of forces F-

(the so-called Kanzaki forces) on lattice sites m f in the ideal


lattice to yield the displacement
si = G^-')^'

(4)

where G_ ^ is now the lattice Green's function.


Asymptotically
s.S = S1(RS=P) , G1^r-RS1 )Fk-

-Pnk3nGik

(S)

with the dipole tensor


p

nk

= Y mp

^n

k~

T61
lbj

corresponding to a force density (r) = F-6(r_-R-) with Xn- the


location in the lattice at which the force is applied.
Exact agreement with the asymptotic result is achieved for
a force density given by
VD = -pn*3*C!>
<7>
which is obtained by going to the limit F- -> , R -> O in such a
way that Pn^ stays constant. As an illustration, the dilatation
centre consisting of three orthogonal double forces of equal
strength is shown in figure 1. This has the force density
f(r) = (fi,f7,f3) with f,(r) = f [6(x1-h)-6(xi+h)]6(x2)6(x3),
^2Tr) = f0[6tx2-h)-6(x2+A)]6(xi)$(x3), and 3tr) = fo[6(x3-h)-6
(x3 + h) ] 6 (x-j_) 6 (x2) with the dipole tensor
"1 O O]
P.. = 2h 6.. = 2 h 0 1 0
U
o ij
o ^Q J
By taking the limit h - O, fo -* (h0, fixed), we obtain the
concentrated force density -j_ = - 2hf O 8 i<$ (r_) = -Pij3j<5(r_) =
-P086(r_) representing an infinitesimal centre of expansion. (We
define P0 = Tr E/3 = Pii/3, where double bars underneath a
symbol denote a tensor quantity).

In general, the dipole tensor of an interstitial defect in


a bcc lattice is expected to be anisotropic rather than isotropic, as shown in Figure 5-1 for the dilatation center. Methods
to calculate the dipole tensor from atomistic models based on
equation (6) have been described (3). An alternate approach (4)
involves a calculation of the change in energy as a function of
applied strain of an atomistic model with and without a point
defect. The first strain derivatives give the various components of the dipole tensor (5) while the second derivatives
provide values for the change in elastic constants or, equivalently, in the dipole tensor on introducing a point defect into
a strained crystal. By analogy with electrostatics, this second
quantity is also called the diaelastic polarizability of the
defect, i.e. it represents the change in the dipole tensor due
to an applied strain.
The total external volume change Av (also called the
relaxation volume of the point defect) of a finite crystal to
whiclr one defect has been added depends on the forces generated
by the defect only through the dipole tensor and the compliances
of the crystal. It is independent of the position of the forces
as long as these are not too close to the free surface, as well

Figure
5-1: Schematic diagram showing an isotropic double
force tensor centered at the origin of the coordinate system
and directed along the coordinate axes X-, , x~ , x^. A center
of dilatation is shown.

as the size of the crystal.

For a cubic crystal

Av - P 1 1 XCc 1 1+ Zc12) - ^i ;

(9)

where the c's are elastic constants and K is the bulk modulus.
Although macroscopically, the defect is completely
described by its dipole tensor, the displacements close to it
must be described in terms of an atomic model. Such a model
permits a detailed local characterization of the defect in terms
of, for instance, the distribution of Kanzaki forces (6). These
are fictitious forces that, when applied to the atoms in the
perfect harmonic lattice, give the same displacements as the
atoms experience in the imperfect lattice. The Kanzaki forces
have no real physical meaning since they include contributions
from changed coupling constants and anharmonicities. Their real
advantage in representing the distortion field of a point defect
is their short range compared to the displacements s. The longrange displacements can be expressed by the dipole tensor which,
in atomic models, can be readily calculated using equation (6).
The relaxation volume is then simply determined from equation
(9). On the other hand, a calculation of Av based directly on
the :s_ determined in an atomic model would require a much larger
modeT size.
It is nevertheless possible and useful to obtain a crude

estimate of the displacements, even close to the point defect,


from continuum theory based on the simple model of a concentrated
force density centered at the defect (determined by P alone), as
illustrated by equations (7) and (8) for the dilatation center.
In a finite crystal, the displacements can be separated into the
sum of two components, a component .s00 appropriate to an infinite
crystal plus a component s* (lEimage), which satisfies the
boundary condition of a stress-free surface. We have
s_ = s + s 1

(10)

Analogously, the total volume change can be written as


Av = Av + Av 1 = (l + Y)Av

y = ^(11)
Av
For the simplest case of a dilatation center at the origin of a
sphere of isotropic elastic material (see equations (7) and (8)

(12)
, R = radius of crystal

(13)

and

(14)

Hence,
(isotropic, elastic)

(15)

I
3 3
where G is the shear modulus. From the ratio s_ /s_ ^ r /R , it
is evident that s_I is negligible near the defect, but that both
displacements are of equal order near the surface where Av is
measured. It should be noted that, for the dilatation center in
an isotropic crystal, also the separate contributions Av00 and
Av 1 are each constant for any size and shape of the crystal or
the position of the forces in it. However, in general, because
the divergence of s may not be concentrated at the defect site,
a simple relationship between Av and Av00, as given by equations
(11) and (15) , cannot be derived, since the separate contributions Av00 and Av^ may depend on the size and shape of the
crystal. In contradistinction, their sum, Av, is always insensitive to these details.
2.2

Interaction Energy

The total potential energy for a material subject to forces


_ (including surface forces) is

(16)

where s_ is an exact solution for a given _.


fully separated into, say, two parts

If f_ can be meaning-

f = a + fb

(17)

then the total energy splits into three parts with the first two
parts giving the self - energies caused by the forces f3- and fb
and the third part giving the interaction energy Ejnt = Ea^ as
follows:

(18)
This can be reduced to the following equivalent expressions

(19)

where Q and g are stress and strain tensors, respectively.


The forces in equation (17) can refer to one defect as
described, for instance, by its double force tensor E and its
interaction with applied surface forces, or with forces due to
another defect such as another point defect, a dislocation or a
crack. The separation given by equation (17) is meaningful as
long as the forces fa and f" do not overlap.
A very simple first-order result comes about if the applied
g-field is uniform, or varies only slowly, over the volume of
the defect centered at r o , which can be simulated by a fairly
concentrated set of source forces. This is the case when the
source forces _f can be written in terms of the defect's dipole
tensor according to equation (7), yielding for the interaction
energy
E.int_ = -P.13.e.ij.v (r
oj )

J
^ (20)

where PJJ = dipole tensor of the defect and e-jj = strain field
acting at the location of the defect (in the absence of the
defect) for given external or internal sources of stress. When
the point defect acts as a dilatation center, equation (20) can
be further simplified to
E.int. = -P o 6 .ij- e -ij- (^ r o)
j

J
^(21)

This shows that a defect that can be described by an isotropic


dipole tensor which interacts only with the hydrostatic component of the applied stress (or strain).
Following the methods of Eshelby (1), the results of this
and the previous section can also be derived based on the strain
representation. This is a continuum model in which the point
defect is represented as a misfitting inclusion, i.e. having
dimensions different from the hole provided for it in the solid.
In this representation, the interaction described by equations
(20) and (21) is often also called the size-effect interaction.
For the dilatation defect, this yields

mt = - ^r-Av

(22)

For an anisotropic point defect, the long-range distortion


field of the point defect is characterized by the X-tensor (7),
which represents the misfit strain of the point defect determined
at the outer boundaries of the crystal. This is related to the
dipole tensor E, using Hooke's law, by
P

ij = c ij,k X la
In terms of ^, the interaction energy is written
E

int ' - a ij X ij

<>

^24a)

or, equivalently, using Eshelby's notation (1)


E

int = -ViJ6Ij1
<24b)
T
where V 5 is the volume of the defect site and the e^j are the
transformation-free strains.
2.3

The Diaelastic Polarizability

The treatment of the interaction energy given by equations


(19) to (24) is incomplete. In addition to the permanent distortions due to the defect, which can be described by the dipole
tensor at zero applied strain (the permanent dipole tensor), the
applied stress can also induce additional strains near the
defect. These arise because the ideal lattice force constants
are usually changed in the vicinity of the defect, resulting in
a different response to an applied field there compared with
elsewhere in the crystal. The defect can also be characterized
by induced Kanzaki forces or, macroscopically, by an induced
dipole force tensor Agd defined as follows:
AP1/ E

Pij(E)

- Pij(0) = aij)k/Ek)l

(25)

The tensor a^^ -^. ^ is called the diaelastic polarizability of


the defect ana is a fourth-rank tensor having the symmetries of
the elastic moduli.
Writing equation (18) in terms of the induced forces 6jfa
and 6_b, we obtain for the induced interaction:
E^nt = - h

dr(6as_b + 6fV)

(26)

This shows that, generally, we must consider not only the forces
induced by one source on the other, but also the reverse
process. For a point defect interacting with a dislocation, a
crack or externally applied loads, the influence of the point
defect on these sources can be neglected. Making use of equation (25), the interaction energy can then be written
E

int ' - * eij(ro>aij,iad ek*<V>

(27)

where |(r ) is the applied strain acting at the defect site r .


Note that the interaction energy caused by the induced dipole is
of second order compared with the interaction caused by the
permanent dipole. The diaelastic polarizability of a defect can
be calculated from atomic models (4) as indicated earlier. It
is related to the change in elastic constants when adding a
small concentration C 1 of defects to a crystal by

6c

ij,ia = - a i j , k /

C28)

(where ti is the atomic volume). In principle, this relation


provides a means of determining the polarizability experimentally.
Again, following the methods of Eshelby (1), the same
result can be obtained in the strain representation.
In this
continuum model, the defect is thought to consist of a region
that is assumed to have "elastic constants" different from those
of the surrounding matrix and will, therefore, distort differently under an applied stress. By expressing the interaction
energy in terms of the observable change in elastic constants on
adding a dilute solution of the point defects to the solid, the
interaction energy is written:

(29)
where W^ and W^ are the interaction energy densities for a
single defect due to, respectively, the dilatational and the
deviatoric part of the induced strain. Equation (29) is often
called the inhomogeneity interaction. More explicitly by
writing

(30)

with

(31)

(where v is Poisson's ratio) we obtain:

(32)

(writing a-jj = a/3 + ' a ij> a = aij) ^he same result could be
obtained by combining equations 1^7) and (28).
2.4

The Paraelastic Polarizability

For a defect having an anisotropic dipole tensor, application of an external stress causes certain orientations of the
defect to have lower energy than others. There is, in general,
an energy barrier between these states, which means that their
accessibility is governed by thermal activation. This gives
rise to a temperature-dependent polarization which, by analogy
with dipoles in electricity and magnetism, is called the paraelastic polarizability. Experimentally, a non-vanishing paraelastic polarizability results in an anelastic relaxation called
the Snoek effect (7) and in differences in direction of expansion of single crystals to which defects have been added.
From equation (21) , the interaction energy of a point

defect having dipole P with an applied external field g depends


on the orientation of E with respect to . By analogy with the
diaelastic polarizability, the paraelastic change APp in the
dipole tensor is defined by (7)
AP

ij P = a ij,ia P e k*

(")

ij,UP = T^T [<PijPk^-<Pij><Pk^]

(34)

where
a

and < > denotes an average over equivalent orientations. Otherwise equivalent orientations of the defect will have different
energies and, in thermal equilibrium, the probability of orientation is proportional to exp(-E- t/kgT). An expansion in
terms of the deviation of the dipole tensor from its isotropic
value, P O ,
F^ = PO + 6E^

(35)

yields the paraelastic polarizability


a i j > k / = I 6P1.* 6Pk//ZkBT

(36)

where Z is the number of orientations. The paraelastic polarizability can be measured as a change in elastic constants according to equation (28). Note that, unlike a^j ^ , it is temperature dependent. Generally, however, the paraelastic values are
about an order of magnitude larger than the diaelastic ones,
except at low temperatures where the orientations of the dipole
may be frozen in.
3.

THERMODYNAMICS OF STRESSED SOLIDS CONTAINING MOBILE


INTERSTITIALS

The solubility of hydrogen in a stressed lattice is determined by its chemical potential. In an inhomogeneously stressed
solid, care must be taken in defining the chemical potential of
a component of the solid, since this may not have a unique
value. As demonstrated by Li, Oriani and Darken (8), for a
mobile component such as hydrogen dissolved interstitially in a
solid subjected to a generalized state of stress, a unique
chemical potential can be defined everywhere. This chemical
potential y^ is given by
^H

where y^
_ 9w
H ~ Sn^
WH

+ W

H -WH

(37)

= chemical potential of hydrogen at zero stress


_ partial molal strain energy of component H(the
change in strain energy of the solid because
of the addition of n^ moles of hydrogen)
= work done per mole addition of hydrogen.

For a dilute solution of hydrogen in iron, the hydrogen

solubility C^ in a stressed solid is, therefore, written


CHQ = CH exp[(WH - wH)/RT]
where Cu

(38)

is the solubility in the unstressed solid.

Assuming uniform applied stress within a small region of


the body, Li, Oriani and Darken (8) show that

H = Vij ^
l
n'

where V is the volume of the body and n' denotes constancy of


moles of the other components of the solid. Equation (39) can
be rewritten in the form
W

H = 0Ij(ViJ

(4)

where (VH)-JJ is a generalization of V H , the partial molal volume


of hydrogen in the solid, assuming isotropic expansion. In
going from partial molal to partial atomic quantities, we see
that, except for a trivial change in in sign, equation (40) is
identical to the size-effect interaction energy of equations
(20) to (24).
The strain energy is given by
w

= Jveijcij,kadeka

(41

which, for the partial molal strain energy, yields


w

ff

u
M v uHc -xj,k*
- 10
H = Jj

+ v

P|_cc/ii
k*lJ ,Jt1 - -ijd e k
8n \
H

(42)

The first term is simply the change in strain energy because of


the added strain contained in the hydrogen atoms while the
second quantity is the change caused by the modification of the
elastic constants resulting from the addition of hydrogen to the
solid. Again, in going from partial molal to partial atomic
quantities, the second term in equation (42) is identical to the
inhomogeneity or diaelastic polarizability term of eouations
(27) to (32).
Generally, the w^-term is an order of magnitude smaller
than wj-[ and can be neglected. As the simplest example, _ in a
sample under uniaxial tensile stress GU, and assuming Vpj is
isotropic, the solubility is given by
CHa = CH exp(anVH/3RT)

(43)

This shows that a tensile stress raises the solubility of hydrogen in the_solid. Measurement of this increase can be used to
determine V^.
4.

APPLICATION TO HYDROGEN IN IRON

The preceding sections have shown that a quantitative description of the elastic interaction of hydrogen in the lattice

of bcc iron requires knowledge of (a) the defect's dipole tensor


(or, equivalently, its strain or partial moIaI volume tensor)
and (b) its diaelastic and paraelastic polarizabilities or
defect moduli tensors.
In turn, these quantities depend on the
interstitial's site occupancy.
In the following, the experimental and theoretical evidence providing information on these
parameters is examined. A recent short review of this has been
given by Hirth (9) .
4.1

Interstitial Site

Hydrogen in iron has a small solubility and is easily


trapped at defect sites. Because of this, there is little
direct evidence about the nature of the sites occupied in the
dissolved state. Impurity interstitials in a bcc lattice are
generally assumed to occupy either octahedral or tetrahedral
sites, as shown in Figures 5-2 (a) and (b).

Figure 5-2: Location of the interstitial in a bcc lattice


showing the surrounding lattice atoms in a) the tetrahedral
configuration, and b) the octahedral configuration.

Each host atom has three octahedral and six tetrahedral sites.
Both of these have tetragonal symmetry. Therefore, one would
expect that the asymptotoc distortion field of hydrogen occupying either of these sites could be described by a tetragonal
dipole tensor which, referred to cube axes, has the general
form:

P (tetragonal) =

[AOO"
O B O
LP O B_

(44)

In general, one should expect that the magnitude of the tetragonality, defined by |A-B|, would depend on the interstitial
site chosen. This will be discussed in the next section.
Although the hydrogen site in iron has not yet been determined directly, evidence from studies of hydrogen in other bcc
metals and from muons in iron favours the tetrahedral site. In
particular, studies using incoherent neutron scattering (10),
(11), channeling (12), diffuse (13) and quasi-elastic neutron
scattering (14) , and neutron spectroscopy involving both
acoustic and optical modes (15) have indicated that, in niobium,
hydrogen occupies the tetrahedral site. Based on spin precession measurements of muons in a-iron (muons can be thought of
as somewhat lighter versions of hydrogen, having one-seventh the
mass), Seeger (16) has concluded that these particles occupy
tetrahedral sites; however, based on the limited available evidence, occupancy of octahedral interstices could not be
completely ruled out.
A thermodynamic analysis of hydrogen solubility in iron by
da Silva et al. (17) attributes the solubility in the 6-iron
temperature range to an increase in entropy, arising from an
increased occupancy of the octahedral, at the expense of the
tetrahedral, sites. This dual-occupancy model predicts mostly
tetrahedral site occupancy in the a-iron temperature range,
which is consistent with the indirect evidence from most other
bcc metals. For a sufficiently anisotropic dipole tensor, there
should also be a Snoek-type internal friction peak, but early
interpretations (18) , (19) that a peak in the 20 to 40 K range
was caused by this have since been revised (9). The difficulty
in obtaining a peak attributable to the Snoek effect may be
caused in part by other factors besides the low hydrogen solubility or its strong trapping tendency. As discussed in the
next section, the tetragonality of the dipole tensor, which also
determines the strength of the signal, may be quite small for
hydrogen in tetrahedral sites.
Recent theoretical calculations (20) indicate that hydrogen
in niobium occupies tetrahedral sites. In contrast, these
calculations predict that positive muons should occupy octahedral sites in niobium, contrary to the conclusions gleaned
from experiment (16). In addition, an examination of the origin
of the relative stability of the tetrahedral versus octahedral
sites leads to the conclusion that a decrease in the size of the
host atom (smaller lattice parameter) acts in the same way as an
increase in the size of the interstitial atom. Both effects
tend to stabilize the octahedral site and can be used to explain
why large interstitial atoms such as carbon, nitrogen and oxygen,
as well as hydrogen in chromium, favour octahedral sites. These
tendencies suggest that hydrogen in iron should also choose
octahedral sites.
4.2

Partial Molal Volume and Strain Field Symmetry

As indicated generally in Section 2.1, introduction of an


interstitial hydrogen atom into iron expands the lattice.
Methods of measuring this volume change caused by dissolved
hydrogen in a metal have been reviewed by Peisl (21).
The most
direct method involves measurement of the changes AV/V in the
external dimensions of the crystal as hydrogen is added. This
yields Av, or equivalently V^, according to

(45)
where ^p6 and Vp e are the atomic and molal volumes of iron,
respectively.
The expansion caused by a single hydrogen atom may be
anisotropic. However, when the hydrogen is randomly distributed
and oriented, the average volume change will be isotropic.
Under these conditions, it is more convenient to measure the
length change, rather than the volume change, of a sample to
which hydrogen has been added. Care must, however, be taken to
ensure that the dimensional changes observed are caused only by
the addition of hydrogen. An alternate procedure, not sensitive
to the change in total sample volume, is to measure the change
in lattice parameter with hydrogen concentration. Additionally,
changes in the solubility caused by an applied stress can,
according to equation (43), be used to obtain an estimate of VHCaution must be exercised in using this relation if the defect
has, as expected, a tetragonal macroscopic strain field. In
this case, the applied stress will cause a preferential ordering
of the hydrogen atoms, which needs to be considered in analyzing
the results. The assumption that the molal volume is isotropic
will, if erroneous, lead in this case to an overestimation of
VH- Along similar lines, measurements of the anelastic strain
caused by the Gorsky effect (the spatial redistribution of hydrogen in a dilatation gradient) can be used to determine a value
for Av (21). The most complete information, capable of providing both the magnitude and symmetry of the distortion field
of hydrogen, can be obtained from scattering experiments such as
Huang diffuse scattering of X-rays and diffuse neutron scattering (21). The effects observed from all these methods depend
linearly on the hydrogen concentration, thus severely limiting
their applicability in iron at low temperatures, where the hydrogen concentration is low. Conversely, where measurements are
possible, an accurate determination of the hydrogen concentration is essential.
Hirth (9) has recently reviewed the existing measurements
of VH in iron. Table 1 summarizes the experimental findings.
An average value of VH = 2.0 x 10~ 6 m 3 mol" 1 (3.3 x 10"3U m^/atom)
emerges from these data. This is close to the empirically
determined "size" of the hydrogen atom of about 3 . O x 10-30 m3/
atom, which has been found in most of the materials in which it
has been measured (21). The degree of tetragonality of the
dipole tensor cannot be measured directly. The result of VH =
2.6 x 10~6m3mo}-l for Armco iron in tension compared with VH =
2.0 x lO'^m^mol"1 from dilatometric measurements is, therefore,
intriguing as it suggests the possibility of a tetragonal dipole
tensor.
Measurements of the dipole tensor of hydrogen in other bcc
metals such as niobium and tantalum, where the tetrahedral site
occupancy has been established, indicate a nearly isotropic
dipole tensor. There have been a number of attempts to explain
this result theoretically. Bauer et al. (13) suggest that an
isotropic dipole tensor can be achieved by a suitable distribution of forces on first- and second-nearest neighbours to the
interstitial hydrogen atom. Birnbaum and Flynn (28) proposed a
motional-averaging model in which hydrogen is assumed to tunnel
between neighbouring interstitial sites. The reported anomalies
in the Debye-Waller factor and the specific heat at low temperatures, which prompted the proposal for this model, however, were

later shown to be caused by, in the case of the former, an


inappropriate analysis of the data (29) and, in the case of the
latter, the presence of other interstitial impurities such as
oxygen and/or nitrogen (30). Thus, there do not appear to be
any experimental data requiring the existence of tunneling
states. Moreover, there seems to be ample evidence, as cited by
Sugimoto and Fukai (20) , demonstrating that hydrogen atoms are
localized in space.
Recent calculations of hydrogen in niobium by Sugimoto and
Fukai (20), based on an, albeit, highly simplified quantum
mechanical model of the hydrogen atom and using a hydrogen-metal
pair potential fitted, following Bauer et al. (13), to give an
almost isotropic dipole tensor for hydrogen in the tetrahedral
sites, show that a large anisotropic (tetragonal) dipole tensor
is obtained for hydrogen in octahedral sites. The order of
magnitude of the tetragonality agrees with measurements of interstitial occupancy of the octahedral site (hydrogen in vanadium
(31), (32), oxygen, nitrogen and carbon in bcc metals (7).
Their conclusions regarding site occupancy have been discussed
in the previous section. The large differences in the tetragonality of the dipole tensor of hydrogen, with respect to tetrahedral and octahedral site occupancy, leads to the interesting
possibility that the presence of a uniaxial stress could
actually stabilize the octahedral sites. Sugimoto and Fukai
(20) cite experimental evidence, from vanadium hydride systems,
that indicate this possibility.
4.3

Defect Moduli

The addition of hydrogen to iron can affect the elastic


moduli in two ways. Firstly, the addition of hydrogen alters
the force constants in the solid. This is a consequence of both
the hydrogen-iron interaction and the overall dilatation of the
solid caused by the hydrogen. Measurements of elastic constant
changes due to these effects provide information on the
diaelastic polarizability of the defect, as given by equations
(27) and (32).
Secondly, the presence of an applied stress can
bias the orientation of a defect with an anisotropic dipole
tensor, giving rise to the Snoek effect. Measurements of the
changes in elastic constants caused by this latter effect provide information on the paraelastic polarizability of the defect
(section 2.4) which in turn relates to the tetragonality of the
dipole tensor.
For an interstitial atom in a tetrahedral site of a bcc
metal, only the modulus c 1 = (c;Q-ci2)/2 is affected, and the
change Ac' is given by (2)
Ac' = ^-^ (A-B)2
oukgi

(46)

where we have assumed a dipole tensor of the form of equation


(44). Equation (46) shows that the change in c 1 is linearly
proportional to the hydrogen concentration C^ and inversely proportional to the temperature T. There is also no effect on the
bulk modulus because no reorientation can occur under dilatation
or pressure, nor on c^^, because of the assumed tetragonality of
the defect (2). It is evident that equation (46) provides a
means of measuring this tetragonality. Note that the relaxation
due to the paraelastic (Snoek) effect can only be detected if
the jump frequencies of the defect are fast enough. For the
experimental methods employed, this does not pose a serious

limitation for a fast diffuser such as hydrogen.


Accurate values of the defect moduli for hydrogen in iron
that are attributable to hydrogen in solution have not been
determined as yet. Analogous results for hydrogen in the Group
VB bcc metals have been attributed to either a small tetragonality (33), (34) or a tetragonality that is smaller than the
accuracy of the measurements (35) , (36). This conclusion by the
latter authors (35), (36) was corroborated by the result that
the measured temperature dependence was not as expected from a
Snoek effect. In contrast, the absence of a 1~
dependence was
deemed by Alberts et al. (34) not to be a sufficient condition
to repudiate the Snoek mechanism because, according to them,
such a dependence has not been observed in other systems where
the Snoek mechanism is otherwise well established. Moreover,
Alberts et al. (34), following the j-ray studies of Blaschko et
al. (37), show that the tetragonality of the hydrogen atom
increases under hydrostatic pressure.
This does not, however,
appear to rule out the possibility that the tetragonality is
negligible at atmospheric pressure, as determined, for instance,
by Huang diffuse X-ray scattering (21).
5.

CONCLUSIONS

(a) Within linear elasticity, general expressions are given


for the energy of a point defect interacting with another strain
or stress field. The relations given are valid when the strain
is due to either internal or external sources. Different microscopic representations of the defect lead to the same interaction energy expression. The interaction energy depends only
on macroscopically determinable quantities of the defect, such
as the force dipole tensor (or the macroscopic strain tensor)
and the defect moduli.
(b) The expansion of a finite crystal due to the addition
of one hydrogen atom is proportional to the trace of the dipole
tensor. For a defect having cubic symmetry, the trace of the
dipole tensor, or, equivalently, its atomic or molal relaxation
volume is sufficient to calculate the dominant term in the
interaction energy.
(c) The magnitudes and symmetries of the dipole and the
defect moduli tensors depend on the location of the hydrogen
atom in the lattice of iron.
(d) Because of its small size, hydrogen dissolved in iron
is expected to enter the lattice interstitially, on either
tetrahedral or octahedral sites. Evidence from other bcc metal
systems indicates tetrahedral site occupancy. No direct confirmation of this exists for the iron-hydrogen system.
(e) Occupation of both the tetrahedral and octahedral sites
in a bcc lattice is expected to result in tetrahedral defect
symmetry. Experimental evidence on impurity interstitials in
metals with bcc crystal structure indicates that occupancy of
the tetrahedral site results in either a very small or a negligible amount of tetragonality. On the other hand, the occupancy
of the octahedral site results in a substantial tetragonality.
In the former case of negligible tetragonality, the dominant
part of the interaction energy is simply given by either of two
equivalent expressions: i) the product of the applied hydrostatic stress times the relaxation volume of the defect in a
finite crystal, or ii) the product of the trace of the defect's
dipole tensor times the applied dilatation.
(f) Experimental determinations of the molal volume of
hydrogen in iron yield an average value of 2.0 x 10"^m^mol"^.

(g) Questions concerning site occupancy, dipole tensor


symmetry and defect moduli of hydrogen in iron, of importance in
evaluating the elastic interaction energy, remain unanswered at
the present time.

Table 1:

Comparison Between Hydrogen Partial Molal Volumes in


Iron Alloys
C

V
H6
(xlO"b

Temp
( C)

H
(atom
fraction)

400-800

>4.7xlO~5

2.1

>.7.3xlO"5

1.6

Armco iron
and steel

20-80

2.0
1.8

Armco iron
AISI 4340
steel

27-60
27-60

2.6
2 .0

Reference

Method

Material

Wagenblast
and Wriedt
(22)

DiIatometry

Ferrovac E

Raczynski
(23)

DiIatometry

High-purity
iron (15
appm)
impurities)

Beck et al .
(24)
de Kazinczy
(25) as cal.
by Oriani
(26)

Effect
of
stress
on
permeability

Bockris et
al. (27)

Effect
of
stress
on
permeabil ity

IU3IIlOl"1)

REFERENCES
1.
2.
3.
4.
5.
6.
7.
8.
9.
10.

Eshelby, J. D., in: Solid State Physics (F. Seitz and D.


Turnbull, eds . ) , Vol. 3, pp. 79-144, Academic Press, New
York (1956) .
Leibfried, G. and Breuer, N., in: Point Defects in Metals I,
Springer, Berlin (1978).
Schober, H. R. and Ingle, K. W., J. Phys . F: Metal Phys .
10: 575-81 (1980).
Dederichs, P. H., Lehmann, C. and Scholz, A., Z . Phys .
B 20: 155-163 (1975) .
Schober, H. R. and Zeller, R., J. Nucl . Mater. 69 70:
341-349 (1978).
Kanzaki, H., J. Phys. Chem. Solids 2: 24-36 (1957).
Nowick, A. S. and Berry, B. S., Anelastic Relaxation in
Crystalline Solids, Academic Press, New York (1972).
Li, J. C. M., Oriani, R. A. and Darken, L. S., Z . Phys .
Chemie Neue Folge 49: 271-290 (1966).
Hirth, J. P., Met. Trans. UA: 861-890 (1980).
Verdan, G., Rubin, R., Kley, W. in Neutron Inelastic
Scattering, pp. 223-31, Proc . IAEA Symp. Copenhagen, Vol.

11.
12.
13.
14.
15.
16.
17.
18.
19.
20.
21.
22.
23.
24.
25.
26.
27.
28.
29.
30.
31.
32.
33.
34.
35.
36.
37.

I, IAEA, Vienna (1968).


Birchall, J. H. L. and Ross, D. D., International Meeting
on Hydrogen in Metals, Julich, March, 1972, Ber.
Kernforschungsanlage Julich 6: 313-45 (1972).
Antonini, M. and Carstanjen, H. D., Phys. Status Solidi (a)
34: K153-157 (1976).
Bauer, G. S., Schmatz, W. and Just, W., Proceedings of the
2nd International Congress on Hydrogen in Metals, Paris
1977, Pergamon, Oxford, 1978, Vol. 2, 2C15.
Stump, N., Gissler, W. and Rubin, R., Phys. Status Solidi
(b) 54: 295-302 (1972) .
Springer, T., in: Hydrogen in Metals I Basic Properties
(G. Alefeld and J. Volkl, eds.), Vol. 28, pp. 75-100,
Springer, Berlin (1978).
Seeger, A., in ref. 15, pp. 349-397.
da Silva, J. R. G., Stafford, S. W. and McLellan Rex B.,
J. Less-common Metals 49: 407-420 (1976).
Heller, W. R., Acta Metall. 9: 600-613 (1961).
Gibala, R., Trans. Met. Soc. AIME 239: 1574-85 (1967).
Sugimoto, H. ancTFukai, Y., Phys. Rev. B 22: 670-680 (1980).
Peisl, J., in ref. 15, pp. 53-75.
Wagenblast, H. and Wriedt, H. A., Met. Trans. 2: 1393-1397
(1971).
Raczynski, W., as quoted in reference 9.
Beck, W., Bockris, J. O'M., McBreen, J. and Nanis, L.,
Proc. R. Soc. London 29OA: 220-235 (1966).
de Kazinczy, F., Jernkonterets Ann. 139: 885-892 (1955).
Oriani, R. A., Trans. Met. Soc. AIME 236: 1368-1369 (1966).
Bockris, J. O'M., Beck, W., Genshaw, M. A., Subramanyan, P.
K. and Williams, F. S., Acta Metall. 19: 1209-1218 (1971).
Birnbaum, H. K. and Flynn, C. P., Phys. Rev. Lett. 37:
25-28 (1976).
Lottner, V., Heim, A. and Springer, T., Z. Phys. B 32: 157165 (1979).
Morkel, C., Wipf, H. and Neumaier, K., Phys. Rev. Lett. 40:
947-950 (1978).
Westlake, D. G., Ockers , S. T., Mueller, M. FI. and Anderson,
K. D., Met. Trans. 3: 1709-1710 (1972).
Asano, H., Abe, Y. and Hirabayashi, M., Acta Metall. 24:
95-99 (1976).
Fisher, E. S., Westlake, D. G. and Ockers, S. T., Phys.
Status Solidi (a) 28: 591-602 (1975).
Alberts, H. L., Fisher, E. S., Katahara, K. W. and
Manghnani, M. H., J. Phys. F: Metal Phys. 9: L209-L213
(1979).
Buchholz, J., Volkl, J. and Alefeld, G., Phys. Rev. Lett.
30: 318-321 (1973).
Magerl, A., Berre, B. and Alefeld, G., Phys. Status Solidi
(a) 36: 161-171 (1976).
Blaschko, 0., Klemencic, R., Weinzierl, P. and Eder, O. J.,
J. Phys. F: Metal Phys. 8: L149-L151 (1978).

Potrebbero piacerti anche